0% found this document useful (0 votes)
509 views268 pages

Students Notes

This document contains lecture notes on wind energy systems. It covers the historical uses of wind power, characteristics of wind such as worldwide distribution and variation with height, aerodynamics of wind turbines including rotor theory, and performance of wind turbines. The notes are intended for students and include topics such as early windmill designs, US wind research in the 1900s, modern turbine designs, atmospheric stability, wind measurement techniques, analysis of wind data, actuator disc theory, blade element theory, momentum theory breakdown, and methods of regulating turbine speed and power output.

Uploaded by

1balamanian
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
509 views268 pages

Students Notes

This document contains lecture notes on wind energy systems. It covers the historical uses of wind power, characteristics of wind such as worldwide distribution and variation with height, aerodynamics of wind turbines including rotor theory, and performance of wind turbines. The notes are intended for students and include topics such as early windmill designs, US wind research in the 1900s, modern turbine designs, atmospheric stability, wind measurement techniques, analysis of wind data, actuator disc theory, blade element theory, momentum theory breakdown, and methods of regulating turbine speed and power output.

Uploaded by

1balamanian
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 268

2013

DIPARTIMENTO DI INGEGNERIA INDUSTRIALE

Pavesi Giorgio

Wind Energy
Systems
NOTES FOR STUDENTS
Summary
SUMMARY ....................................................................................................................................... I
1 NOMENCLATURE................................................................................................................ VI
1.1 ACRONYMS ...................................................................................................................... VII
2 LIST ...................................................................................................................................... VIII
2.1 FIGURES .......................................................................................................................... VIII
2.2 TABLE .............................................................................................................................. XV
3 HISTORICAL USES OF WIND ......................................................................................... - 1 -
3.1 HISTORICAL DEVELOPMENT ........................................................................................... - 1 -
3.2 FIRST UNITED STATES RESEARCHES .............................................................................. - 5 -
3.3 OTHER RESEARCHES .................................................................................................... - 12 -
3.4 DARRIEUS WIND TURBINES.......................................................................................... - 16 -
3.5 INNOVATIVE WIND TURBINES ...................................................................................... - 18 -
3.6 CURRENT STATUS AND FUTURE PROSPECTS .................................................................. - 22 -
3.7 REFERENCE .................................................................................................................. - 25 -
4 WIND CHARACTERISTICS ........................................................................................... - 27 -
4.1 METEOROLOGY OF WIND ............................................................................................. - 27 -
4.2 WORLD DISTRIBUTION OF WIND .................................................................................. - 42 -
4.2.1 Other atmospheric circulation patterns ................................................................................. - 45 -
4.2.2 Variations in time ................................................................................................................... - 47 -
4.2.3 Variation due to wind direction ............................................................................................. - 50 -
4.3 ATMOSPHERIC STABILITY .......................................................................................... - 51 -
4.4 WIND SPEED VARIATION WITH HEIGHT ....................................................................... - 58 -
4.4.1 Turbulence ............................................................................................................................. - 60 -
4.4.2 Acceleration effect ................................................................................................................. - 60 -
4.5 MEASUREMENT OF WIND .............................................................................................. - 62 -
4.5.1 Ecological indicators ............................................................................................................. - 62 -
4.5.2 Anemometers .......................................................................................................................... - 63 -
4.6 ANALYSIS OF WIND DATA - 68 -
4.6.1 Average wind speed ............................................................................................................... - 68 -
4.6.2 Weibull Statistics .................................................................................................................... - 71 -
4.6.3 Rayleigh and Normal Distributions ....................................................................................... - 80 -

| Summary Pavesi i
4.6.4 Distribution of Extreme Winds ............................................................................................... - 87 -
4.7 REFERENCE .................................................................................................................. - 93 -
5 AERODYNAMICS OF WIND TURBINES..................................................................... - 95 -
5.1 THE ACTUATOR DISC CONCEPT ................................................................................... - 96 -
5.2 THE BETZ LIMIT ........................................................................................ - 100 -
5.3 ROTOR DISC THEORY ................................................................................................. - 104 -
Maximum power ................................................................................................................................... - 106 -
5.4 WAKE STRUCTURE ........................................................................................ - 108 -
5.5 SHROUTED ROTORS .................................................................................................... - 109 -
5.6 VORTEX CYLINDER MODEL OF THE ACTUATOR DISC ................................................ - 112 -
5.7 AXIAL FLOW FIELD ........................................................................................ - 121 -
5.8 TANGENTIAL FLOW FIELD ........................................................................................... - 121 -
5.9 RADIAL FLOW FIELD ................................................................................................... - 122 -
5.10 CONCLUSIONS ............................................................................................................ - 123 -
6 ROTOR BLADE THEORY ......................................................................................... - 124 -
6.1 2-D BLADE AERODYNAMICS ...................................................................................... - 124 -
6.2 BLADE ELEMENT THEORY ........................................................................................ - 130 -
6.3 DETERMINATION OF ROTOR TORQUE AND POWER ....................................................... - 133 -
7 BREAKDOWN OF THE MOMENTUM THEORY .................................................... - 136 -
7.1 FREE-STREAM/WAKE MIXING ..................................................................................... - 136 -
7.2 MODIFICATION OF ROTOR THRUST CAUSED BY FLOW SEPARATION ............................. - 136 -
7.3 EMPIRICAL DETERMINATION OF THRUST COEFFICIENT ............................................... - 137 -
7.4 THE EFFECTS OF A DISCRETE NUMBER OF BLADES .................................................... - 138 -
7.4.1 Tip-losses ............................................................................................................................. - 138 -
8 WIND –TURBINE PERFORMANCE ........................................................................... - 143 -
8.1 THE CP –  PERFORMANCE CURVE .............................................................................. - 143 -
8.1.1 The effect of solidity on performance ................................................................................... - 143 -
8.1.2 The CM –  curve.................................................................................................................. - 145 -
8.1.3 The CT –  curve .................................................................................................................. - 145 -
8.2 CONSTANT ROTATIONAL SPEED OPERATION.............................................................. - 146 -
8.2.1 The KP – 1/ curve ............................................................................................................... - 146 -
8.2.2 Stall regulation .................................................................................................................... - 146 -
8.2.3 Effect of rotational speed change ......................................................................................... - 147 -

| Summary Pavesi ii
8.2.4 Effect of blade pitch angle change ....................................................................................... - 148 -
8.2.5 Pitch regulation ................................................................................................................... - 148 -
8.2.6 Pitching to stall .................................................................................................................... - 148 -
8.2.7 Pitching to feather................................................................................................................ - 149 -
8.3 VARIABLE-SPEED OPERATION .................................................................................... - 149 -
8.4 REFERENCE ................................................................................................................ - 149 -
9 POWER CONTROL ........................................................................................................ - 151 -
9.1.1 Energy Capture .................................................................................................................... - 151 -
9.1.2 Mechanical Loads ................................................................................................................ - 152 -
9.1.3 Power Quality ...................................................................................................................... - 152 -
9.2 MODES OF OPERATION ............................................................................................... - 153 -
9.3 CONTROL STRATEGIES ............................................................................................... - 154 -
9.3.1 Fixed-speed Fixed-pitch....................................................................................................... - 154 -
9.3.2 Fixed-speed Variable-pitch .................................................................................................. - 157 -
9.3.3 Variable-speed Fixed-pitch .................................................................................................. - 166 -
9.3.4 Variable-speed Variable-pitch ............................................................................................. - 168 -
9.4 YAW CONTROL ........................................................................................................... - 169 -
9.5 BRAKING SYSTEMS..................................................................................................... - 169 -
9.5.1 Independent braking systems—requirements of standards .................................................. - 169 -
9.5.2 Active pitch control .............................................................................................................. - 170 -
9.5.3 Pitching blade tips ............................................................................................................... - 170 -
9.5.4 Spoilers ................................................................................................................................ - 171 -
9.5.5 Other devices ....................................................................................................................... - 171 -
9.6 MECHANICAL BRAKE OPTIONS ................................................................................... - 172 -
9.6.1 Parking versus idling ........................................................................................................... - 172 -
9.7 FIXED SPEED, TWO-SPEED OR VARIABLE-SPEED ........................................................ - 172 -
9.7.1 Two-speed operation ............................................................................................................ - 173 -
9.7.2 Variable-speed operation..................................................................................................... - 173 -
9.7.3 Variable-slip operation ........................................................................................................ - 179 -
9.7.4 Other approaches to variable-speed operation ................................................................... - 180 -
9.8 REFERENCES............................................................................................................... - 181 -
10 PERFORMANCE OF WIND ENERGY CONVERSION SYSTEMS ........................ - 183 -
10.1 POWER CURVE OF THE WIND TURBINE ........................................................................ - 183 -
10.2 ENERGY GENERATED BY THE WIND TURBINE .............................................................. - 186 -

| Summary Pavesi iii


10.2.1 Weibull based approach ...................................................................................................... - 186 -
10.2.2 Rayleigh based approach ..................................................................................................... - 189 -
10.3 CAPACITY FACTOR ...................................................................................... - 190 -
10.4 MATCHING THE TURBINE WITH WIND REGIME ............................................................ - 191 -
10.5 PERFORMANCE OF WIND POWERED PUMPING SYSTEMS ............................................... - 194 -
10.5.1 Wind driven piston pumps .................................................................................................... - 194 -
10.5.2 Rotor performance ............................................................................................................... - 194 -
10.5.3 Wind pump characteristics................................................................................................... - 197 -
10.5.4 Integrated system performance ............................................................................................ - 197 -
10.5.5 Wind driven roto-dynamic pumps ........................................................................................ - 198 -
10.5.6 Integration of wind regime characteristics .......................................................................... - 200 -
10.5.7 Wind electric pumping systems ............................................................................................ - 200 -
10.6 REFERENCES............................................................................................................... - 202 -
11 ENVIRONMENTAL ASPECTS AND IMPACTS ........................................................ - 203 -
11.1 LIFE CYCLE ANALYSIS ................................................................................................ - 204 -
11.1.1 Net energy analysis .............................................................................................................. - 207 -
11.1.2 Life cycle emission ............................................................................................................... - 210 -
11.2 AVIAN INTERACTION WITH WIND TURBINES .............................................................. - 212 -
11.2.1 Overview of the problem ...................................................................................................... - 212 -
11.2.2 Characterization of the problem .......................................................................................... - 212 -
11.3 CURRENT MITIGATION CONCEPTS AND CASE STUDIES................................................. - 213 -
11.3.1 Mitigation concepts .............................................................................................................. - 213 -
11.3.2 Case studies ......................................................................................................................... - 214 -
11.4 RESOURCES FOR ENVIRONMENTAL ASSESSMENT STUDIES OF AVIAN IMPACTS ............ - 214 -
11.5 VISUAL IMPACT OF WIND TURBINES .......................................................................... - 219 -
11.5.1 Characterization of the problem .......................................................................................... - 219 -
11.5.2 Design of wind systems to minimize visual impact .............................................................. - 221 -
11.5.3 Visual impact mitigation design strategies for US sites ....................................................... - 221 -
11.5.4 European wind farm visual impact design characteristics .................................................. - 222 -
11.5.5 Resources for visual impact studies ..................................................................................... - 222 -
11.5.6 US visual impact resources .................................................................................................. - 223 -
11.5.7 European visual impact resources ....................................................................................... - 223 -
11.6 WIND TURBINE NOISE ...................................................................................... - 223 -
11.6.1 Noise and sound fundamentals ............................................................................................ - 224 -
11.6.2 Characteristics of sound and noise ...................................................................................... - 224 -

| Summary Pavesi iv
11.6.3 Sound power and pressure measurement scales .................................................................. - 225 -
11.6.4 Measurement of sound or noise ........................................................................................... - 226 -
11.6.5 Noise mechanisms of wind turbines ..................................................................................... - 227 -
11.6.6 Aerodynamic noise ............................................................................................................... - 228 -
11.6.7 Mechanical noise ................................................................................................................. - 229 -
11.6.8 Noise prediction from wind turbines .................................................................................... - 230 -
11.6.9 Single wind turbines ............................................................................................................. - 230 -
11.6.10 Multiple wind turbines ......................................................................................................... - 231 -
11.6.11 Noise propagation from wind turbines ................................................................................ - 232 -
11.6.12 Noise reduction methods for wind turbines ......................................................................... - 233 -
11.6.13 Noise standards or regulations ............................................................................................ - 233 -
11.7 ELECTROMAGNETIC INTERFERENCE EFFECTS............................................................. - 234 -
11.7.1 Characterization of electromagnetic interference from wind turbines ................................ - 235 -
11.7.2 Mechanism of electromagnetic interference ........................................................................ - 235 -
11.7.3 Wind turbine parameters ..................................................................................................... - 236 -
11.7.4 Potential electromagnetic interference effects of wind turbines .......................................... - 237 -
11.7.5 Prediction of electromagnetic interference effects from wind turbines: analytical models . - 238 -
11.7.6 Simplified analysis ............................................................................................................... - 238 -
11.7.7 Resources for estimation ...................................................................................................... - 240 -
11.8 LAND-USE ENVIRONMENTAL IMPACTS ........................................................................ - 242 -
11.8.1 Land-use considerations ...................................................................................................... - 242 -
11.8.2 Mitigation of land-use problems .......................................................................................... - 243 -
11.9 OTHER ENVIRONMENTAL CONSIDERATIONS .............................................................. - 244 -
11.9.1 Safety.................................................................................................................................... - 244 -
11.9.2 Mitigation of public health safety risks ................................................................................ - 245 -
11.9.3 Impact on flora and fauna .................................................................................................... - 246 -
11.9.4 Mitigation measures............................................................................................................. - 247 -
11.9.5 Shadow flicker and pashing ................................................................................................. - 247 -
11.9.6 Mitigation measures............................................................................................................. - 248 -
11.10 REFERENCES............................................................................................................... - 249 -

| Summary Pavesi v
1 Nomenclature

| Nomenclature Pavesi vi
1.1 Acronyms

| Nomenclature Pavesi vii


2 List
2.1 Figures
Figure 3.1.1 Illustration of ancient Persian wind mill. ..................................................................... - 1 -
Figure 3.1.2 An ancient windmill in the British Isles (Author: Michael Reeve, source:
Wikipedia, http://wikipedia.org. GNU Free Documentation License applies to
this image) .................................................................................................................... - 2 -
Figure 3.1.3 An ancient Spanish ‘wind farm’ (Author: Lourdes Cardenal, source: Wikipedia,
http://wikipedia.org. GNU Free Documentation License applies to this image) ......... - 3 -
Figure 3.1.4 Typical American multiblade windmill. ....................................................................... - 4 -
Figure 3.2.1 NSF–NASA MOD-0 wind power system: (a) general view; (b) superstructure
and equipment. Rated power output, 100 kW; rated wind speed, 8 m/s (18 mi/h). ...... - 6 -
Figure 3.2.2 MOD-0A located at Kahuku Point, Oahu, Hawaii. ..................................................... - 7 -
Figure 3.2.3 The MOD OA wind turbine (Courtesy of Hawaiian Electric Company, Inc.,
http://heco.com) ............................................................................................................ - 7 -
Figure 3.2.4 MOD-1 located at Boone, North Carolina. (Courtesy of DOE.) ................................. - 9 -
Figure 3.2.5 MOD-2 located at the Goodnoe Hills site near Goldendale, Washington. .................. - 9 -
Figure 3.3.1 1.5 MW, 64 m diameter Wind Turbine ....................................................................... - 12 -
Figure 3.3.2 750 kW, 48 m diameter Wind Turbine in Denmark .................................................... - 12 -
Figure 3.3.3 1 MW Wind Turbine in Northern Ireland ................................................................... - 13 -
Figure 3.3.4 Wind Farm of Variable-Speed Wind Turbines in Complex Terrain .......................... - 13 -
Figure 3.3.5 Wind Farm of Six Pitch-regulated Wind Turbines in Flat Terrain ............................ - 14 -
Table 3.3.1 Installed Wind Turbine Capacity Throughout the World, January 2001................... - 15 -
Figure 3.4.1 Sandia Laboratories 17-m Darrieus, rated at 60 kW in a 12.5-m/s wind. ................. - 16 -
Figure 3.4.2 Extruded aluminium blade of 17-m Darrieus............................................................. - 16 -
Figure 3.5.1 Kansas State University Savonius, rated at 5 kW in a 12-m/s wind. .......................... - 18 -
Figure 3.5.2 Typical performances of wind machines. ................................................................... - 19 -
Figure 3.5.3 Magnus force on a spinning cylinder. ........................................................................ - 20 -
Figure 3.5.4 Madaras concept for generating electricity. .............................................................. - 20 -
Figure 3.5.5 Augmented vortex turbine. ......................................................................................... - 21 -
Figure 4.1.1 Wind Spectrum Farm Brookhaven Based on Work by van der Hoven (1957) ........... - 29 -
Figure 4.1.2 Schematic energy cycle without anthropogenic interference. The energy flows
are in TW (1012W). Numbers in parentheses are uncertain or rounded off (with
use of Odum, 1972; Budyko, 1974; Gregg, 1973; Lorenz, 1967; Manabe, 1969).
The earth receives around 1.7 1014 kW of power from the sun in the form of
solar radiation. This radiation heats up the atmospheric air. ................................... - 30 -
Figure 4.1.3 Coordinates in a geocentric picture. P represents the North Pole, Z zenith and
φ the geographical latitude. The unit radius great circle shown through P and
Z is the median of the observer. S represents the direction to the Sun, and the
solar coordinates are the declination δ and the hour angle ω. Also shown is the

| List Pavesi viii


zenith angle z (equal to π/2 minus the height of the Sun over the horizon). ............... - 31 -
Figure 4.1.4 Daily average radiation, E SW 0 (Wm-2), incident at the top of the Earth’s
atmosphere, as a function of time of the year and latitude (independent of
longitude). Latitude is taken as positive towards north (in this and the following
figures) (based on NCEP-NCAR, 1998). .................................................................... - 31 -
Figure 4.1.5 a, b Top of the atmosphere albedo for the Earth atmosphere system, derived
from satellite data for January (top) and July (bottom) 1997 (NCEP-NCAR,
1998). .......................................................................................................................... - 32 -
Figure 4.1.6 Latitude dependence (averaged over longitudes and over the year) of incident (
0  ) and outgoing long-wavelength (  E 0  ) radiation at the top of the
E sw sw

atmosphere, the amount of radiation absorbed (A0) and reflected (R0) by the
Earth− atmosphere system, and the net radiation flux E0 at the top of the
atmosphere (in Wm-2; based on Sellers, 1965; Budyko, 1974). ................................. - 32 -
Figure 4.1.7 The poleward energy transfer required by the latitude variation of the net
energy flux shown in Figure 4.1.6. The dashed curve represents the same
quantity divided by the length of the latitude circle, i.e. the energy transport
required across a unit length of the latitude circle. These figures are annual
and longitudinal averages (based on Sellers, 1965). The ocean transport
contribution is indicated as a dark blurred region, based on measurements with
estimated uncertainties (Trenberth and Solomon, 1994). .......................................... - 34 -
Figure 4.1.8 Altitude dependence of temperature, pressure and density for standard
atmosphere. ................................................................................................................ - 34 -
Figure 4.1.9 a,b Average temperature in the atmosphere as a function of height and latitude,
for January (above) and July (below) 1995, both at longitude zero (based on
NCEP-NCAR, 1998). .................................................................................................. - 36 -
Figure 4.1.10 Geotrophic wind ......................................................................................................... - 36 -
Figure 4.1.11 Coriolis force .............................................................................................................. - 38 -
Figure 4.1.12 Wind forces in a low-pressure area ........................................................................... - 38 -
Figure 4.1.13 Wind forces in a high-pressure area .......................................................................... - 38 -
Figure 4.1.14 Illustration of the geostrophic wind; Fp, pressure force on the air; Fp, Coriolis
force ............................................................................................................................ - 39 -
Figure 4.1.15 Illustration of the gradient wind; Ugr R, radius of curvature ..................................... - 39 -
Figure 4.2.1 Terrestrial pressure and wind systems ....................................................................... - 42 -
Figure 4.2.2 a, b Zonal winds (at longitude zero) for January and April of 1997, as functions
of latitude and pressure (height) (based on NCEP-NCAR, 1998). ............................ - 43 -
Figure 4.2.3 a, b Zonal winds (at longitude zero) for July and October of 1997, as functions
of latitude and pressure (height) (based on NCEP-NCAR, 1998). ............................ - 43 -
Figure 4.2.4 Summer (top) and winter (bottom) streamfunctions for the meridional mass
transport (longitude average), given in units of 109 kg s-2. Cells with a positive
streamfunction has northward transport at top (based on Newell et al., 1969). ....... - 44 -
Figure 4.2.5 Northward transport of angular momentum by eddies, averaged over year and
longitudes and expressed as 1018 kg m2s-1 per unit of pressure layer, i.e. per

| List Pavesi ix
104 N m-2. Expressed per unit of height z, the magnitude of transport would
increase less rapidly upwards towards the two maxima (based on Buch, 1954;
Obasi, 1963; also using Lorenz, 1967). ..................................................................... - 44 -
Figure 4.2.6 Diurnal valley and mountain wind ............................................................................. - 46 -
Figure 4.2.7 Time and space scales of atmospheric motion (Spera, 1994). ................................... - 46 -
Figure 4.2.8 Seasonal changes of monthly average wind speeds (Hiester and Pennell, 1981) ..... - 47 -
Figure 4.2.9 Seasonal variation of available wind power per unit area (Rohatgi and Nelson,
1994). .......................................................................................................................... - 48 -
Figure 4.2.10 Monthly mean diurnal wind speeds for January and July (Hiester and Pennell,
1981) ........................................................................................................................... - 48 -
Figure 4.2.11 Typical plot of wind speed vs. time for a short period ............................................... - 49 -
Figure 4.2.12 Illustration of a discrete gust event; a, amplitude; b, rise time; c, maximum gust
variation; d, lapse time ............................................................................................... - 49 -
Figure 4.2.13 Time series of monthly wind speeds (Hiester and Pennell, 1981).............................. - 49 -
Figure 4.3.1 Pressure variation with altitude for U.S. Standard Atmosphere ................................ - 51 -
Figure 4.3.2 Dry-adiabatic and actual temperature variations with height as a function of
time of day, for clear skies.......................................................................................... - 53 -
Figure 4.3.3 Buoyancy of air in a stable atmosphere ..................................................................... - 54 -
Figure 4.3.4 Deep buoyant layer topped by temperature inversion. .............................................. - 54 -
Figure 4.3.5 Western Kansas wind data, 1970-1973. ..................................................................... - 56 -
Figure 4.4.1 Variation of wind velocity with height ....................................................................... - 58 -
Figure 4.4.2 Velocity ratio with respect to 10 m for different roughness heights .......................... - 58 -
Figure 4.4.3 Variation of wind profile exponent α with reference wind speed u(z1). .................... - 60 -
Figure 4.4.4 Turbulence created by an obstruction ........................................................................ - 61 -
Figure 4.4.5 The acceleration effect over ridges ............................................................................ - 61 -
Figure 4.5.1 Biological rating scales for the wind speed ............................................................... - 62 -
Figure 4.5.2 Cup anemometer......................................................................................................... - 63 -
Figure 4.5.3 Principle of pressure plate anemometer .................................................................... - 64 -
Figure 4.5.4 Principle of pressure tube anemometer...................................................................... - 64 -
Figure 4.5.5 Sonic anemometer ...................................................................................................... - 65 -
Figure 4.5.6 Vanes to measure wind direction ............................................................................... - 67 -
Figure 4.5.7 Wind roses showing the distribution of frequency, velocity and energy in
different directions ............................................................ Error! Bookmark not defined.
Figure 4.6.1 General relationship between (a) a density function f(u) and (b) a distribution
function F(u) ............................................................................................................... - 70 -
Figure 4.6.2 Weibull density function f(u) for scale parameter c = 1. ........................................... - 72 -
Figure 4.6.3 Weibull scale parameter c divided by mean wind speed u versus Weibull shape
parameter k................................................................................................................. - 73 -
Figure 4.6.4 Cumulative distribution generated using different methods ...................................... - 77 -
Figure 4.6.5 Actual wind data, and Weibull and Rayleigh density functions for Tuttle Creek,
7:00–8:00 a.m., August 1980, 10 m. .......................................................................... - 81 -

| List Pavesi x
Figure 4.6.6 Actual wind data, and Weibull and Rayleigh density functions for Tuttle Creek,
7:00–8:00 p.m., July 1980, 50 m. ............................................................................... - 81 -
Figure 4.6.7 Confidence intervals for two measured monthly means, u m1 and u m 2 . ................... - 84 -
Figure 4.7.1 The Energy Extracting Stream-tube of a Wind Turbine ............................................. - 95 -
Figure 5.1.1 The Energy Extracting Stream-tube of a Wind Turbine ............................................. - 96 -
Figure 5.1.2 Alternative control volume around a wind turbine .................................................... - 98 -
Figure 5.2.1 Variation of CP, CT and UW/U with Axial Induction Factor a............................... - 100 -
Figure 5.2.2 The measured thrust coefficient CT as a function of the axial induction factor a
and the corresponding rotor states .......................................................................... - 101 -
Figure 5.2.3 Schematic view of the turbulent-wake state induced by the unstable shear flow
at the edge of the wake ............................................................................................. - 101 -
Figure 5.2.4 Schematic view of the kinetic energy transfer by an actuator disk. ......................... - 103 -
Figure 5.3.1 The Trajectory of an Air Particle Passing Through the Rotor Disc ........................ - 104 -
Figure 5.3.2 Tangential Velocity Grows Across the Disc Thickness ............................................ - 104 -
Figure 5.5.1 Ideal flow through a wind turbine in a diffuser........................................................ - 109 -
Figure 5.5.2 Computed mass flow ratio plotted against the computed power coefficient ratio
for a bare and a shrouded wind turbine, respectively .............................................. - 110 -
Figure 5.5.3 Computed power coefficient for a rotor in a diffuser as a function of the thrust
coefficient CT ........................................................................................................... - 110 -
Figure 5.6.1 Prandtl’s induction velocities. .................................................................................. - 112 -
Figure 5.6.2 Flow visualisation with smoke, revealing the tip vortices........................................ - 112 -
Figure 5.6.3 Flow visualisation with smoke, revealing smoke trails being ‘sucked’ into the
vortex spirals. ........................................................................................................... - 113 -
Figure 5.6.4 Helical Vortex Wake Shed by Rotor with Three Blades Each with Uniform
Circulation  .......................................................................................................... - 113 -
Figure 5.6.5 Simplified Helical Vortex Wake Ignoring Wake Expansion ..................................... - 113 -
Figure 5.6.6 The Geometry of the Vorticity in the Cylinder Surface ............................................ - 114 -
Figure 5.6.7 Axial induction factor ............................................................................................... - 118 -
Figure 5.6.8 - 118 -
Figure 5.6.9 Theoretical maximum power coefficient as a function of tip speed ratio for an
ideal horizontal axis wind turbine, with and without wake rotation ........................ - 120 -
Figure 5.7.1 The Radial and Axial Variation of Axial Velocity in the Vicinity of an Actuator
Disc, a=1/3 ............................................................................................................... - 121 -
Figure 5.8.1 The Axial Variation of Tangential Velocity in the Vicinity of an Actuator Disc at
50% Radius, a =1/3, =6 ......................................................................................... - 121 -
Figure 5.8.2 The Axial Variation of Tangential Velocity in the Vicinity of an Actuator Disc at
101 percent Radius, a =1/3, =6 ............................................................................. - 122 -
Figure 5.9.1 Flow Field Through an Actuator Disc for a=1/3 ..................................................... - 122 -
Figure 6.1.1 Schematic view of streamlines past an airfoil .......................................................... - 124 -
Figure 6.1.2 Definition of lift and drag ......................................................................................... - 124 -
Figure 6.1.3 Airfoil characteristics as function of the angle of attack. ........................................ - 125 -

| List Pavesi xi
Figure 6.1.4 Polar for the FX67-K-170 airfoil ............................................................................. - 125 -
Figure 6.1.5 Different stall behaviour .......................................................................................... - 126 -
Figure 6.1.6 Reynold influence on the FX67-K-170 airfoil Lift Coefficient ................................. - 126 -
Figure 6.1.7 Qualitative representation of separation types. ....................................................... - 127 -
Figure 6.1.8 Different stall types for NACA airfoils at Reynolds 5.5·106 [Hoerner, S.F. and
Borst, H.V. (1985)]. .................................................................................................. - 128 -
Figure 6.1.9 Diagrammatic representation of the boundary layer flow near the separation
point [51].................................................................................................................. - 128 -
Figure 6.2.1 A Blade Element Sweeps Out an Annular Ring ........................................................ - 130 -
Figure 6.2.2 Blade Element Velocities and Forces ....................................................................... - 130 -
Figure 6.3.1 Power Coefficient-Tip Speed Ratio Performance Curve ......................................... - 134 -
Figure 6.3.3 Typical variations of Torque Coeffi cient for a variable-pitch wind
turbine ...................................................................................................................... - 134 -
Figure 6.3.4 Typical variations of CM and CP for a fixed-pitch wind turbine. ............................. - 135 -
Figure 7.2.1 Comparison of Theoretical and Measured Values of CT ......................................... - 137 -
Figure 7.4.1 Helical Trailing Tip Vortices of a Horizontal Axis Turbine Wake ........................... - 138 -
Figure 7.4.2 Azithimuthal Variation of a for Various Radial Positions for a Three-blade
Rotor with Uniform Blade Circulation Operating at a Tip Speed Ratio of 6. The
blades are at 120°, 240° and 360°. .......................................................................... - 139 -
Figure 7.4.3 Span-wise Variation of the Tip-loss Factor for a Blade with Uniform
Circulation................................................................................................................ - 139 -
Figure 7.4.4 Span-wise Variation of Power Extraction in the Presence of Tip-loss for a
Blade with Uniform Circulation on a Three-blade Turbine Operating at a Tip
Speed Ratio of 6 ........................................................................................................ - 141 -
Figure 7.4.5 A Helicoidal Vortex Sheet Wake for a Two-blade Rotor.......................................... - 141 -
Figure 8.1.1 CP –  Performance Curve for a Modern Three-blade Turbine Showing Losses. ... - 143 -
Figure 8.1.2 Effect of Changing Solidity....................................................................................... - 144 -
Figure 8.1.3 The Effect of Solidity on Torque ............................................................................... - 144 -
Figure 8.1.4 The Effect of Solidity on Thrust ................................................................................ - 145 -
Figure 8.2.1 Non-dimensional Performance Curves for Constant Speed Operation ................... - 146 -
Figure 8.2.2 Effect on Extracted Power of Rotational Speed ....................................................... - 147 -
Figure 8.2.3 Effect on Extracted Power of Rotational Speed at Low Wind Speed ....................... - 147 -
Figure 8.2.4 Effect on Extracted Power of Blade Pitch Set Angle................................................ - 148 -
Figure 8.2.5 Pitching to Feather Power Regulation Requires Large Changes of Pitch Angle .... - 148 -
Figure 8.4.1 Ideal Power Curve.................................................................................................... - 151 -
Figure 9.2.1 Operating points for different operating conditions ................................................ - 153 -
Figure 9.3.1 Fixed-Speed Fixed Pitch control strategy. ............................................................... - 155 -
Figure 9.3.4 Power Curves for Different Pitch Angles: 40 m Diameter Rotor Rotating at 33
r.p.m.......................................................................................................................... - 158 -
Figure 9.3.5 Pitch Linkage System Used in Conjunction with a Single Hydraulic Actuator
Located in the Nacelle. (The central triangular ‘spider’ is connected to the

| List Pavesi xii


actuator by a rod passing through the hollow low-speed shaft. Links from the
spider drive the blade pitch via braced arms cantilevering into the hub from
each blade. Each arm is parallel to its blade axis, but eccentric to it). ................... - 159 -
Figure 9.3.6 Blade Pitching System Using Separate Hydraulic Actuators for Each Blade.
(Each actuator cylinder is supported on gimbal-type mountings bolted to the
hub, and its piston applies a pitching torque to the blade via a cantilevered
conical tube eccentric to the blade axis. The blade is attached to the outer ring
of the pitch bearing). ................................................................................................ - 160 -
Figure 9.3.7 Blade Pitching System Using a Separate Electric Motor for Each Blade. (A
pinion, driven by the motor via a planetary gearbox, engages with gear teeth on
the inside of the inner ring of the pitch bearing, to which the blade is bolted.
The blade is not attached to the bearing in this photograph, so the fixing holes
are visible). ............................................................................................................... - 161 -
Figure 9.3.10 Specimen Pitch Angle Schedules for Active Pitch Control and Active Stall
Control...................................................................................................................... - 164 -
Figure 9.3.11 Power Curves for Different Negative Pitch Angles: 40 m Diameter Rotor
Rotating at 33 r.p.m. ................................................................................................. - 164 -
Figure 9.3.13 Basic fixed-speed pitch-to-feather and pitch-to-stall control strategies: (a)
output power and (b) power efficiency vs. wind speed ............................................. - 165 -
Figure 9.6. 1 Passive Control of Tip Blade, Using Screw on Tip Shaft and Spring ...................... - 171 -
Figure 9.7.1 Rotor efficiency and annual energy production versus rotor tip-speed ratio. ......... - 174 -
Figure 9.7.2 Shaft power output of Sandia 17-m Darrieus in variable-speed operation. ............ - 175 -
Figure 9.7.3 Shaft torque output of Sandia 17-m Darrieus in variable-speed operation. ............ - 176 -
Figure 9.7.4 Load with square-law torque variation connected to Sandia 17-m turbine............. - 177 -
Figure 9.7.5 Three loads attached to Sandia 17-m turbine .......................................................... - 178 -
Figure 9.7.6 Torque versus rotational speed for low-solidity 10-m-diameter two-bladed
propeller. .................................................................................................................. - 179 -
Figure 9.7.7 Torque versus rotational speed for Savonius turbine: radius 0.88 m; area 22
m2. ............................................................................................................................. - 179 -
Figure 9.7.8 Speed control methods used in small to medium size turbines................................. - 180 -
Figure 10.1.1 Ideal power curve of a pitch controlled wind turbine .............................................. - 184 -
Figure 10.1.2 Field performance of a 250 kW wind turbine........................................................... - 185 -
Figure 10.1.3 Correlation between measured and generated performances ................................. - 185 -
Figure 10.2.1 Graphical representation of energy estimation........................................................ - 187 -
Figure 10.4.1 Effect of cut-out velocity on the system performance. ............................................. - 192 -
Figure 10.4.2 Effect of cut-in velocity on the system performance.................................................. - 192 -
Figure 10.4.3 Comparison of power curves for different rated wind speeds................................. - 193 -
Figure 10.4.4 Capacity factor as a function of rated wind velocity............................................... - 193 -
Figure 10.5.1 Typical power curve of a multi-bladed mechanical wind pump ............................. - 196 -
Figure 10.5.2 Field performance of the experimental rotor .......................................................... - 196 -
Figure 10.5.3 Field performance of wind driven piston pumps ..................................................... - 199 -
Figure 11.1.1 Various stages of the life cycle of a wind turbine..................................................... - 205 -

| List Pavesi xiii


Figure 11.1.2 Various stages involved in LSA ................................................................................ - 206 -
Figure 11.1.3 Energy Payback Ratio of various energy sources. ................................................... - 209 -
Figure 11.1.4 Energy payback period of various energy sources .................................................. - 209 -
Figure 11.1.5 Comparison of life cycle emissions from various energy sources ............................ - 211 -
Figure 11.4.1 Eight categories of wind turbines in the Altamont pass (Orloff and Flannery,
1992) ......................................................................................................................... - 217 -
Figure 11.5.1 Summary of fundamental design principles significant to wind farm
development (Stanton, 1994). ................................................................................... - 221 -
Figure 11.6.1 Wind turbine noise assessment factors (Hubbard and Shepherd, 1990) .................. - 226 -
Figure 11.6.2 Definition of A, B, and C frequency weighting scales (Beranek and Ver, 1992) ..... - 226 -
Figure 11.6.3 Schematic of flow around a rotor blade. (Wagner et al., 1996); U, wind
speed. ........................................................................................................................ - 228 -
Figure 11.6.4 Components and total sound power level for wind turbine (Wagner et al.,
1996); LWA predicted sound power level; ah, airborne; slb, structure borne. ......... - 230 -
Table 13. 11 Typical inputs for Class 2 and 3 noise prediction models Cj Vagner et al.,
1996). ........................................................................................................................ - 232 -
Figure 11.7.1 Scattering of electromagnetic signals by a wind turbine (Wagner et al., 1996). ..... - 235 -
Figure 11.7.2 Model configuration for electromagnetic interference due to a wind turbine
(Sengupta and Senior, 1994). ................................................................................... - 239 -
Figure 11.7.3 Geometry for interference of the electromagnetic path by an obstacle (Van Kats
and Van Rees, 1989). ................................................................................................ - 241 -
Figure 11.9.1 Diagram of shadow flicker calculation (European Commission, 1998); A, B
howes ........................................................................................................................ - 248 -

| List Pavesi xiv


2.2 Table
Table 3.2.1 Specifications of ERDA and DOE .............................................................................. - 10 -
Table 3.3.1 Installed Wind Turbine Capacity Throughout the World, January 2001................... - 15 -
Table 3.6.1 Global leaders in wind energy generation 2012 ........................................................ - 22 -
Table 3.6.2 Wind energy generation ............................................................................................ - 23 -
Table 3.6.3 EU leaders in wind energy generation ....................................................................... - 24 -
Table 4.6.1 Normal Cumulative Distribution Function ................................................................ - 83 -
Table 4.6.2 Probability of finding the variable within q standard deviations of u . ................... - 83 -
Table 4.6.3 Average yearly speeds corrected to 30 m (100 ft) ...................................................... - 86 -
Table 4.6.4 α and β for seven Kansas sites, daily fastest minute in mi/h 50 m 30 m 10 m ........... - 89 -
Table 4.6.5 α and β for five Kansas NWS Stations, daily fastest mile. .......................................... - 90 -
Table 4.6.6 α and β for Topeka and Dodge City using monthly fastest mile for 1949–1968. ....... - 90 -
Table 4.6.7 Days per year a given wind speed or greater is observed. ........................................ - 90 -
Table 9. 1 Advantages of Fixed and Variable Speed Systems ................................................... - 175 -
Table 11. 1 Performance regions of a wind turbine .................................................................... - 184 -
Table 13. 1 Stack emissions of coal, gas, and wind power plants (kg/MWh) ............................. - 203 -
Table 13. 2 Annual reduction in atmospheric emissions due to a 5 MW wind farm ................... - 203 -
Table 13. 3 Materials required for a 600 kW land based wind turbine ...................................... - 206 -
Table 13. 4 Additional materials required for the 600 kW offshore turbine ............................... - 206 -
Table 13. 5 Energy equivalent of different materials used in a 600 kW land based turbine....... - 207 -
Table 13. 6 Energy requirement of additional materials used in a 600 kW offshore turbine ..... - 207 -
Table 13. 7 Emissions from different materials used for a 600 kW land based wind turbine. .... - 210 -
Table 13. 8 Emissions from additional materials required for the offshore wind turbine .......... - 210 -
Table 13. 9 Emissions due to the disposal of materials .............................................................. - 210 -
Table 13. 10 Wind turbine aerodynamic noise mechanisms (Wagner et al., 1996). ..................... - 228 -

| List Pavesi xv
3 Historical Uses of Wind
3.1 Historical Development
Human efforts to harness wind for energy date back to the ancient times for at least 3000 years,
when he used sails to propel ships and boats. Later, wind energy served the humankind by energising
his grain grinding mills and water pumps. During its transformation from these crude and heavy
devices to today’s efficient and sophisticated machines, the technology went through various phases of
development.
There is disagreement on the origin of the concept of using wind for mechanical power. Some
believe that the concept originated in ancient Babylonia. The Babylonian emperor Hammurabi planned
to use wind power for his ambitious irrigation project during seventeenth century B.C. [Golding E
(1976)]. Hero of Alexandria, who lived in the third century B.C., described a simple horizontal axis
wind turbine with four sails which was used to blow an organ [Golding E (1976)].
Others argue that the birth place of wind mills is India. In Arthasastra, a classic work in Sanskrit
written by Kautiliya during 4th century B.C., references are seen on lifting water with contrivances
operated by wind [Sorensen B (1995)]. However, there are no records to prove that these concepts got
transformed to real hardware.
The earliest documented design of wind mill dates back to 200 B.C. The Persians used wind mills
for grinding grains during this period. Those were vertical axis machines having sails made with
bundles of reeds or wood. The grinding stone was
attached to the vertical shaft. The sails were
attached to the central shaft using horizontal
struts. The size of the sails was decided by the
materials used for its fabrication, usually 5 m long
and 9 m tall (Figure 3.1.1).
By the 13th century, grain grinding mills
were popular in most of Europe. The French
adopted this technology by 1105 A.D. and the
English by 1191 A.D. In contrast with the vertical
axis Persian design, European mills had
horizontal axis. These post mills were built with
beautiful structures. The tower mill had a fixed
supporting tower with a rotatable cap which
carried the wind rotor. The tower was usually
built of brick in a cylindrical shape, but was
sometimes built of wood, and polygonal in cross
section. In one style, the cap had a support or tail
extending out and down to ground level. A circle
of posts surrounded the tower where the support
touched the ground. The miller would check the
direction of the prevailing wind and rotate the cap
and rotor into the wind with a winch attached
Figure 3.1.1 Illustration of ancient Persian wind between the tail and one of the posts. The tail
mill. would then be tied to a post to hold the rotor in

| Historical Uses of Wind Pavesi -1-


the proper direction. This process
would be repeated when the wind
direction changed. Protection from
high winds was accomplished by
turning the rotor out of the wind or
by removing the canvas covering
the rotor latticework.
The Dutch, with renowned
designer Jan Adriaenszoon, were the
pioneers in making these mills.
They made many improvements in
the design and invented several
types of mills. Examples are the
tjasker and smock mills. The rotors
were made with crude airfoil profile
to improve the efficiency. Apart
from grain grinding, wind mills
were employed to drain marshy
lands in Holland. The optimization
of the rotor shape probably took a
long time to accomplish. It is
interesting to note that the rotors on
many of the Dutch mills are twisted
and tapered in the same way as
modern rotors and appear to have
nearly optimized the aerodynamic
parameters necessary for maximum
efficiency. The rotors presently on
the tower mills probably do not date
Figure 3.1.2 An ancient windmill in the British Isles back to the original construction of
(Author: Michael Reeve, source: Wikipedia, the tower, but still indicate high
http://wikipedia.org. GNU Free Documentation License quality aerodynamic engineering of
applies to this image) a period much earlier than the
present. These wind mills reached
America by mid-1700, through the Dutch settlers.
This is followed by the water pumping wind mill, which is still considered as one of the most
successful application of wind power. The so-called American multi bladed wind turbine appeared in
the wind energy history by the mid-1800 (Figure 3.1.4).
Relatively smaller rotors, ranging from one to several meters in diameter, were used for this
application. The primary motive was to pump water from a few meters below the surface for
agricultural uses. These water pumpers, with its metallic blades and better engineering design, offered
good field performance. Over six million of such units were installed in US alone, between 1850 and
1930.
The era of wind electric generators began close to 1900’s. The first modern wind turbine,
specifically designed for electricity generation, was constructed in Denmark in 1890. It supplied

| Historical Uses of Wind Pavesi -2-


electricity to the
rural areas. By
1910, several
hundred units with
capacities of 5 to
25 kW were in
operation in
Denmark. During
the same period, a
large wind electric
generator having
17 m ‘picket fence’
rotor was built by
Brush in
Cleveland, Ohio
and the research
undertaken by
LaCour in
Denmark. For the
first time, a speed-
up gear box was
introduced in the
design. This system
operated for 20
years generating its
rated power of 12
kW DC. However,
for much of the
twentieth century
there was little
interest in using
wind energy other
than for battery
charging for
remote dwellings
and these low-
power systems
were quickly
replaced once
access to the
Figure 3.1.3 An ancient Spanish ‘wind farm’ (Author: Lourdes Cardenal, electricity grid
source: Wikipedia, http://wikipedia.org. GNU Free Documentation License became available.
applies to this image) More
systematic methods
were adopted for the engineering design of turbines during this period. With low-solidity rotors and

| Historical Uses of Wind Pavesi -3-


aerodynamically designed blades, these
systems could give impressive field
performance. By 1910, several
hundreds of such machines were
supplying electrical power to the
villages in Denmark. By about 1925,
wind electric generators became
commercially available in the American
market. Similarly, two and three bladed
propeller turbines ranging from 0.2 to 3
kW in capacity were available. The
most common brands were Wincharger
(200 to 1200 W) and Jacobs (1.5 to 3
kW). These were used on farms to
charge storage batteries that were then
used to operate radios, lights, and small
appliances with voltage ratings of 12,
32, or 110 volts. A good selection of 32
Vdc appliances was developed by
industry to meet this demand. Then the
Figure 3.1.4 Typical American multiblade windmill. Rural Electric Administration (REA)
was established by Congress in 1936.
Turbines with bigger capacity were also developed during this period. The first utility-scale
system was installed in Russia in 1931. A 100 kW turbine was installed on the Caspian Sea shore,
which worked for two years and generated about 20,000 kW electricity. Experimental wind plants were
subsequently constructed in other countries like United States, Denmark, France, Germany, and Great
Britain. A significant development in large-scale systems was the 1250 kW Smith–Putnam wind
turbine fabricated by Palmer C. Putman. The turbine was put to use in 1941 at the Grandpa’s Knob,
near Rutland, Vermont [Putnam PC (1948)]. This remarkable machine had a steel rotor 53 m in
diameter, full-span pitch control and flapping blades to reduce loads and was mounted on a 34 m tall
tower. The rotor had a chord (the distance from the leading to the trailing edge) of 3.45 m. Each of the
two blades was made with stainless steel ribs covered by a stainless steel skin and weighed 7300 kg.
The blade pitch (the angle at which the blade passes through the air) was adjustable to maintain a
constant rotor speed of 28.7 r/min. The machine operated for 1100 hours during the next five years, i.e.,
till a blade spar failed catastrophically in 1945, it remained the largest wind turbine constructed for
some 40 years. The project is considered to be a success as it could demonstrate the technical feasibility
of large- scale wind-electric generation.
Some interesting designs of wind turbine were experimented during this period. Darrieus G.J.M,
a French engineer, put forth the design of Darrieus turbine in 1920, which was patented in United Sates
in 1931 [Ramler JR, Donovan RM (1979)]. In contrast with the popular horizontal axis rotors, Darrieus
turbines had narrow curved blades rotating about its vertical axis. During the same period, Julius D.
Madaras invented a turbine working on Magnus effect. Magnus effect is basically derived from the
force on a spinning cylinder placed in a stream of air. Another significant development at this time was
the Savonius rotor in Finland, invented by S.J. Savonius. This rotor was made with two halves of a
cylinder split longitudinally and arranged radially on a vertical shaft. The transverse cross-section of

| Historical Uses of Wind Pavesi -4-


the rotor resembled the letter ‘S’ [Savonius SJ (1931)]. The rotor was driven by the difference in drag
forces acting on its concave and convex halves, facing the wind.
Intensive research on the behaviour of wind turbines occurred during 1950's. The concept of high
tip speed ratio-low solidity turbines got introduced during this period. For example, light-weight
constant-speed rotors were developed in Germany in 1968. They had fibre glass blades attached to
simple hollow towers supported by guy ropes. The largest of this breed was of 15 m diameter with a
rated output of 100 kW. Golding (1955) and Shepherd and Divone in Spera (1994) provide a
fascinating history of early wind turbine development. They record the 100 kW 30 m diameter
Balaclava wind turbine in the then USSR in 1931 and the Andrea Enfield 100 kW 24 m diameter
pneumatic design constructed in the UK in the early 1950s. In this turbine hollow blades, open at the
tip, were used to draw air up through the tower where another turbine drove the generator. In Denmark
the 200 kW 24 m diameter Gedser machine was built in 1956 while Electricité´ de France tested a 1.1
MW 35 m diameter turbine in 1963. In Germany, Professor Hutter constructed a number of innovative,
lightweight turbines in the 1950s and 1960s. In spite of these technical advances and the enthusiasm,
among others, of Golding at the Electrical Research Association in the UK there was little sustained
interest in wind generation until the price of oil rose dramatically in 1973.
In the later years, cheaper and more reliable electricity, generated from fossil fuel based plants
became available. When the electricity generated from wind costed 12 to 30 cents/kWh in 1940, the
same generated from other sources was available at 3 to 6 cents/kWh [Kloeffler RG, Sitz EL (1946)].
Cost of electricity from fossil fuels further declined below 3 cents/kWh by 1970. Fossil fuels were
available in plenty at a relatively cheaper rate at that time. Several nuclear power projects were also
embarked on, believing that it would be the ultimate source for the future energy needs. Thus, the
interest in wind energy declined gradually, especially by 1970.
The oil crisis in 1973, however, forced the scientists, engineers and policy makers to have a
second thought on the fossil fuel dependence. They realised that political tampering can restrict the
availability and escalate the cost of fossil fuels. Moreover, it was realised that the fossil fuel reserve
would be exhausted one day or the other.
Nuclear power was unacceptable to many, due to safety reasons. These factors caused the revival
of interest in wind energy. Research on resource analysis, hardware development, and cost reduction
techniques were intensified. The sudden increase in the price of oil stimulated a number of substantial
Government-funded programmes of research, development and demonstration.

3.2 First United States Researches


In United States the Federal Wind Energy Program had its beginning in 1972 when a joint Solar
Energy Panel of the National Science Foundation (NSF) and the National Aeronautics and Space
Administration (NASA) recommended that wind energy be developed to broaden the Nation’s energy
options for new energy sources.[ Ramler, J. R. and R. M. Donovan (1979)]. In 1973, NSF was given
the responsibility for the Federal Solar Energy Program, of which wind energy was a part. The Lewis
Research Center, a Federal Laboratory controlled by NASA, was selected to manage the technology
development and initial deployment of large wind turbines. Early in 1974, NASA was funded by NSF
to:
 design, build, and operate a wind turbine for research purposes, designated the MOD-0,
 initiate studies of wind turbines for utility application, and
 undertake a program of supporting research and technology development for wind turbines.

| Historical Uses of Wind Pavesi -5-


In 1975, the responsibility within the Federal government for wind turbine development was
assigned to the newly created Energy Research and Development Administration (ERDA). ERDA was
then absorbed by the
Department of Energy (DOE)
in 1977. The NASA Lewis
Research Center continued to
direct the technology
development of large turbines
during this period.
During the period
following 1973, other Federal
Laboratories became involved
with other aspects of Wind
Energy Collection Systems
(WECS). Sandia Laboratories,
a DOE Laboratory located at
Albuquerque, New Mexico,
became responsible for
federally sponsored research
on Darrieus wind turbines.
Battelle Pacific Northwest
Laboratories, Richland,
Washington, became
responsible for wind resource
assessments. Solar Energy
Research Institute, (now the
National Renewable Energy
Laboratory) Golden, Colorado,
became responsible for
innovative wind turbines.
Small wind turbine research
was handled by Rockwell,
International at their Rocky
Flats plant near Golden,
Colorado. Agricultural
applications were handled by
the U. S.
Department of
Agriculture from facilities at
Beltsville, Maryland, and
Bushland, Texas. This division
of effort allowed existing
Figure 3.2.1 NSF–NASA MOD-0 wind power system: (a) general personnel and facilities to be
view; (b) superstructure and equipment. Rated power output, 100 kW; shifted over to wind power
rated wind speed, 8 m/s (18 mi/h). research so that results could

| Historical Uses of Wind Pavesi -6-


be obtained in a relatively short time.
It was decided very early in the program that the MOD-0 would be rated at 100 kW and have a
38-m-diameter rotor with two blades [Third Wind Energy Workshop]. This machine would incorporate
the many advances in aerodynamics,
materials, controls, and data handling
made since the days of the Smith-
Putnam machine. The choice of the
two bladed propeller over some more
unusual wind turbines was made on
the basis of technology development.
The two bladed machines had been
built in larger sizes and had been
operated more hours than any other
type, hence had the highest
probability of working reasonably
well from the start. For political
reasons it was important to get
something working as soon as
possible. This machine became
Figure 3.2.2 MOD-0A located at Kahuku Point, Oahu, Hawaii. operational in September, 1975, at
the NASA Plumbrook facility near
Sandusky, Ohio.
A diagram of the turbine and the contents
of the nacelle (the structure or housing on top of
the tower which contains the gearbox, generator,
and controls) is shown in Figure 3.2.1. The rotor
and nacelle sit on top of a 4-legged steel truss
tower about 30 m high. The rotor is downwind
of the tower, so the wind strikes the tower before
striking the rotor. Each rotor blade thus sees a
change in wind speed once per revolution when
it passes through the tower shadow. This
introduces vibration to the blades, which has to
be carefully considered in blade design.
An upwind design tends to introduce
vibration in the tower because of blade
shadowing so neither design has strong
advantages over the other. In fact, the MOD-0
was operated for brief periods as an upwind
machine to assess some of the effects of upwind
operation on structural loads and machine
control requirements.
The MOD-0 was designed so the rotor
Figure 3.2.3 The MOD OA wind turbine would turn at a constant 40 r/min except when
(Courtesy of Hawaiian Electric Company, Inc., starting up or shutting down. A gear box
http://heco.com)

| Historical Uses of Wind Pavesi -7-


increases the rotational speed to 1800 r/min to drive a synchronous generator which is connected to the
utility network. Startup is accomplished by activating a control which aligns the wind turbine with the
wind. The blades are then pitched by a hydraulic control at a programmed rate and the rotor speed is
brought to about 40 r/min.
At this time an automatic synchronizer is activated and the wind turbine is synchronized with the
utility network. If the wind speed drops below the value necessary to get power from the rotor at 40
r/min, the generator is disconnected from the utility grid, the blades are feathered (pitched so no power
output is possible) and the rotor is allowed to coast to a stop. All the steps of startup, synchronization,
power control, and shutdown are automatically controlled by a microprocessor.
The stresses in the aluminium blades were too high when the unit was first placed into operation,
and it was determined that the tower shadow was excessive. The tower was blocking the airflow much
more than had been expected. A stairway inside the tower which had been added late in the design was
removed and this solved the problem.
Except for this tower blockage problem, the MOD-0 performed reasonably well, and provided a
good base of experience for designing larger and better turbines. The decision was made in 1975 to
build several of these turbines, designated as the MOD-0A. The size of tower and rotor remained the
same, but the generator size was doubled from 100 to 200 kW. The extra power would be produced in
somewhat higher wind speeds than the rated wind speed of the MOD-0. The Westinghouse Electric
Corporation of Pittsburgh, Pennsylvania was the prime contractor responsible for assembly and
installation. The blades were built by the Lockheed California Company of Burbank, California. The
first MOD-0A was installed at Clayton, New Mexico in late 1977, the second at Culebra, Puerto Rico
in mid 1978, the third at Block Island, Rhode Island in early 1979, and the fourth at Kahuku Point,
Oahu, Hawaii in early 1980. The first three machines used aluminium blades while the Kahuku MOD-
0A used wood composite blades. The wooden blades weighed 1360 kg each, 320 kg more than the
aluminium blades, but the expected life was longer than their aluminium counterparts. A MOD-0A is
shown in Figure 3.2.2.
The Kahuku machine is located in a trade wind environment where relatively steady, high speed
winds are experienced for long periods of time. The machine produced an average power output of 178
kW for the first 573 hours of operation. This was an outstanding record compared with the output of
the other MOD-0A machines of 117 kW at Culebra, 89 kW at Clayton, and only 52 kW at Block Island
during the first few months of operation for these machines. This shows the importance of good site
selection in the economical application of large wind turbines.
Following the MOD-0 and MOD-0A was a series of other machines, the MOD-1, MOD-2, etc.
Design parameters for several of these machines are shown in Table 3.2.1. The MOD-1 was built as a
2000-kW machine with a rotor diameter of 61 m. It is pictured in Figure 3.2.4. Full span pitch control
was used to control the rotor speed at a constant 35 r/min. It was built at Howard’s Knob, near Boone,
North Carolina in late 1978. It may be noticed from Table 3.2.1 that the rated windspeed for the MOD-
1 was 14.6 m/s at hub height, significantly higher than the others. This allowed the MOD-1 to have a
rated power of 10 times that of the MOD-0A with a swept area only 2.65 times as large.

| Historical Uses of Wind Pavesi -8-


The gearbox and generator were similar in
design to those of the MOD-0A, except larger.
The tower was a steel, tubular truss design.
The General Electric Company, Space Division, of
Philadelphia, Pennsylvania was the prime
contractor for designing, fabricating, and installing
the MOD-1. The Boeing Engineering and
Construction Company of Seattle, Washington,
manufactured the two steel blades.
As theMOD-1 design effort progressed, it
became clear that the MOD-1 would be relatively
heavy and costly and could not lead to a cost
competitive production unit. Weight and cost were
being determined by a number of factors, the most
significant of which were the stiff tower design
criteria, the full span pitch control which required
complicated, heavy mechanisms and excessive
Figure 3.2.4 MOD-1 located at Boone, North
space in the hub area, and a heavy bedplate
Carolina. (Courtesy of DOE.)
supporting the weight on top of the tower. A
number of possible improvements in the design
became evident too late to be included in the actual
construction. Only one machine was built because
of the predicted production costs. Like the MOD-0,
it was operated as a test unit to help the designs of
later generation turbines.
One early problem with the MOD-1 was the
production of sub-audible vibrations which would
rattle the windows of nearby houses. The rotor
would interact with the tower to produce two pulses
per revolution, which resulted in a vibration
frequency of about 1.2 Hz. Techniques used to
reduce the annoyance included reducing the speed
of rotation and replacing the steel blades with
fibreglass blades. Other operational problems,
including a broken low speed shaft, plus a reduction
in federal funding, caused the MOD-1 to be
disassembled in 1982. The next machine in the
series, the MOD-2, represented an effort to build a
truly cost competitive machine, incorporating all
the information gained from the earlier machines. It
Figure 3.2.5 MOD-2 located at the Goodnoe was a second generation machine with the Boeing
Hills site near Goldendale, Washington. Engineering and Construction Company serving as
the prime contractor. The rotor had two blades, was
91.5 m in diameter, and was upwind of the tower. Rotor speed was controlled at a constant 17.5 r/min.
Rated power was 2500 kW (2.5 MW) at a wind speed of 12.4 m/s measured at the hub height of 61 m.

| Historical Uses of Wind Pavesi -9-


In order to simplify the configuration and achieve a lower weight and cost, partial span pitch control
was used rather than full span pitch control. That is, only the outer 30 percent of the span was rotated or
pitched to control rotor speed and power. This construction feature can be seen in Figure 3.2.5.
To reduce the loads on the system caused by wind gusts and wind shear, the rotor was designed to
allow teeter of up to 5 degrees in and out of the plane of rotation. These load reductions saved weight
and therefore cost in the rotor, nacelle, and tower. The word teeter is also used for the motion of a
plank balanced in the middle and ridden by children so one end of the plank goes up while the other
end goes down. This described the same type of motion in the rotor except that motion was around a
horizontal pivot point rather than the vertical one used on the playground.
The MOD-2 tower was designed to be soft or flexible rather than stiff or rigid. The softness of the
tower refers to the first mode natural frequency of the tower in bending relative to the operating
frequency of the system. For a two-bladed rotor, the tower is excited (receives a pulse) twice per
revolution of the rotor. If the resonant frequency of the tower is greater than the exciting frequency, the
tower is considered stiff. A tower is considered soft if the resonant frequency is less than the exciting
frequency, and very soft if the resonant frequency is less than half the exciting frequency. The tower of
the MOD-2 was excited at its resonant frequency for short time periods during startup and shutdown,
so extreme care had to be taken during these times so the oscillations did not build up enough to
damage the tower.
The MOD-2 tower was a welded steel cylindrical shell design. This design was more cost
effective than the stiff, open-truss tower of the first generation machines. The MOD-2 was significantly
larger than the MOD-1, yet the above ground mass was less, 273,000 kg as compared with 290,000 kg.
The first installation of the MOD-2 was a three machine cluster at the Goodnoe Hills site near
Goldendale, Washington, built in early 1981. Two additional units were built, one in Wyoming and one
in California.
The numbering system hit some difficulties at this point, since the next machine after the MOD-2
was the MOD-5. Actually, this third generation machine was designed by two different companies,
with the General Electric version being named the MOD-5A while the Boeing version was named the
Table 3.2.1 Specifications of ERDA and DOE
Two-Bladed Horizontal-Axis Wind Turbines
MOD-0 MOD-0A MOD-1 MOD-2
Rotor r/min 40 40 34.7 17.5
Generator output power (kW) 100 200 2000 2500
Rotor coefficient of performance, CP max 0.375 0.375 0.375 0.382
Cut-in wind speed at hub height (m/s) 4.3 5.4 7.0 6.3
Rated wind speed at hub height (m/s) 7.7 9.7 14.6 12.4
Shutdown wind speed at hub height (m/s) 17.9 17.9 19.0 20.1
Maximum wind speed (m/s) 66 67 66 66
Rotor diameter (m) 37.5 37.5 61 91.5
Hub height (m) 30 30 46 61
Coning angle 7° 7° 12° 0°
Effective swept area (m2) 1072 1140 2920 6560
Airfoil section, NACA- 23,000 23,000 44XX 230 XX
Weight of two blades (kg) 2090 2090 16,400 33,200
Generator voltage, line to line 480 480 4160 4160

| Historical Uses of Wind Pavesi - 10 -


MOD-5B. Objectives of the MOD-2 and MOD-5 programs were essentially identical except that the
target price of electricity was reduced by 25 percent, to 3.75 cents per kWh in 1980 dollars.
The General Electric MOD-5A design called for a rotor diameter of 122 m (400 ft) and a rated
power of 6.2 MW. Rated power would be reached in wind speeds of 13 m/s (29 mi/h) at the hub height
of 76 m (250 ft). The wood rotor would turn at two rotational speeds, 13 or 17 r/min, depending on
wind conditions.
The Boeing MOD-5B in 1987 [Johnson GL (2001)] was designed to be an even larger machine,
7.2 MW with a rotor diameter of 128 m (420 ft). The rotor was designed to be built of steel with wood
tips. A variable speed generator was selected as oppo sed to the fixed speed generator used on the
MOD-2.
Federal research on the MOD series of turbines was terminated in the mid 1980s, and all the
turbines have been scraped. One reason was that smaller turbines (in the 100-kW range) could be built
at lower costs and with better performance than the large turbines. Many of us underestimated the
difficulty of building large reliable wind turbines. The technology step was just too large.
A second reason was that the American aerospace industry did not have a desire to produce a cost
effective commercial product. Wind turbine research was viewed as just another government contract.
A given company would build a turbine on a cost plus basis. When it broke, it would be repaired on a
cost plus basis. When the federal money ran out, the company’s interest in wind power vanished.
Hindsight indicates it would have been far better to have spent the federal money on the small, mostly
undercapitalized, companies that were dedicated to producing a quality wind turbine.
A third reason for the lack of interest in wind was the abundance and depressed costs of
petroleum products throughout the 1980s and into the 1990s. In the mid-1970s, it was standard wisdom
that we were running out of natural gas. Many utilities converted from burning natural gas as a boiler
fuel, instead using coal or nuclear energy. The price of natural gas increased substantially from its
artificially low values. But by the mid-1980s, it was discovered that we had substantial reserves of
natural gas (at this higher price), and utilities started converting back to natural gas as a fuel, especially
for peaking gas turbines. The development of wind power has certainly been delayed by these various
actions of the government, aerospace, and oil industries.

| Historical Uses of Wind Pavesi - 11 -


3.3 Other Researches
Similar programmes were pursued in the UK, Germany and Sweden. There was considerable
uncertainty as to which architecture might prove most cost-effective and several innovative concepts
were investigated at full scale. In Canada, a 4 MW vertical-axis Darrieus wind turbine was constructed
and this concept was also investigated in the 34 m diameter Sandia Vertical Axis Test Facility in the
USA. In the UK, an alternative vertical-axis design using straight blades to give an ‘H’ type rotor was
proposed by Dr Peter Musgrove and a 500 kW prototype constructed. In 1981 an innovative horizontal-
axis 3 MW wind turbine was built and tested in the USA. This used hydraulic transmission and, as an
alternative to a yaw drive, the entire structure was orientated into the wind. The best choice for the
number of blades remained unclear for some while and large turbines were constructed with one, two
or three blades.
Much important scientific and engineering information was gained from these Government-
funded research programmes and the prototypes generally worked as designed. However, it has to be
recognized that the problems of operating very
large wind turbines, unmanned and in difficult
wind climates were often under-estimated and the
reliability of the prototypes was not good. At the
same time as the multi-megawatt prototypes were
being constructed private companies, often with

Figure 3.3.1 1.5 MW, 64 m diameter Wind Figure 3.3.2 750 kW, 48 m diameter Wind
Turbine Turbine in Denmark

| Historical Uses of Wind Pavesi - 12 -


considerable state support, were constructing
much smaller, often simpler, turbines for
commercial sale. In particular the financial
support mechanisms in California in the mid-
1980s resulted in the installation of a very large
number of quite small (100 kW) wind turbines. A
number of these designs also suffered from
various problems but, being smaller, they were in
general easier to repair and modify.
The so-called ‘Danish’ wind turbine
concept emerged of a three-bladed, stall-regulated
rotor and a fixed-speed, induction machine drive
train. This deceptively simple architecture has
proved to be remarkably successful and has now
been implemented on turbines as large as 60 m in
diameter and at ratings of 1.5 MW. The machines
of Figure 3.3.1 and Figure 3.3.2 are examples of
this design. However, as the size of commercially
available turbines now approach that of the large
prototypes of the 1980s it is interesting to see that
the concepts investigated then of variable-speed
operation, full-span control of the blades, and
advanced materials are being used increasingly by
designers.
Figure 3.3.4 shows a wind farm of direct-
Figure 3.3.3 1 MW Wind Turbine in Northern
drive, variable speed wind turbines. In this design,
Ireland
the synchronous generator is coupled directly to
the aerodynamic rotor so eliminating the requirement for a gearbox. Figure 3.3.3 shows a more
conventional, variable-speed wind turbine that uses a gearbox, while a small wind farm of pitch-
regulated wind turbines, where full-span control of the blades is used to regulate power, is shown in

Figure 3.3.4 Wind Farm of Variable-Speed Wind Turbines in Complex Terrain

| Historical Uses of Wind Pavesi - 13 -


Figure 3.3.5.
These projects were stopped by mid-1980’s due to various reasons. During the same period,
scientists at Sandia Laboratories focussed their research on the design and development of the Darrieus
turbine [Sheldahl RE, Blackwell BF (1977)]. They fabricated several models of the Darrieus machine
in different sizes during 1980’s.
Research and development on wind energy are seen intensified in the later years. A few
innovative concepts like the vortex turbine, diffuser augmented design, Musgrove rotors etc. were also
proposed during that time. Prototypes of these turbines were constructed and tested. However, only the
horizontal axis propeller design could emerge successfully on a commercial scale.
All modern electricity-generating wind turbines use the lift force derived from the blades to drive
the rotor. A high rotational speed of the rotor is desirable in order to reduce the gearbox ratio required
and this leads to low solidity rotors (the ratio of blade area/rotor swept area). The low solidity rotor acts
as an effective energy concentrator and as a result the energy recovery period of a wind turbine, on a
good site, is less than 1 year, i.e., the energy used to manufacture and install the wind turbine is
recovered within its first year of operation (Musgrove in Freris, 1990).
The stimulus for the development of wind energy in 1973 was the price of oil and concern over
limited fossil-fuel resources. Now, of course, the main driver for use of wind turbines to generate
electrical power is the very
low CO2 emissions (over
the entire life cycle of
manufacture, installation,
operation and de-
commissioning) and the
potential of wind energy to
help limit climate change.
In 1997 the Commission
of the European Union
published its White Paper
(CEU, 1997) calling for 12
percent of the gross energy
demand of the European
Union to be contributed
from renewables by 2010.
Wind energy was
identified as having a key
role to play in the supply
of renewable energy with
an increase in installed
wind turbine capacity from
2.5 GW in 1995 to 40 GW
by 2010. This target is
likely to be achievable,
there was some 12 GW of
installed wind-turbine
Figure 3.3.5 Wind Farm of Six Pitch-regulated Wind Turbines in Flat capacity in Europe, 2.5
Terrain

| Historical Uses of Wind Pavesi - 14 -


GW of which was constructed in 2000 compared with only 300 MW in 1993. The average annual
growth rate of the installation of wind turbines in Europe from 1993–9 was approximately 40 percent
(Zervos, 2000). The distribution of wind-turbine capacity is interesting with, in 2000,
Germany accounting for some 45 percent of the
Table 3.3.1 Installed Wind Turbine
European total, and Denmark and Spain each having
Capacity Throughout the World, January
approximately 18 percent. There is some 2.5 GW of
2001
capacity installed in the USA of which 65 percent is in
California although with increasing interest in Texas
and some states of the Midwest. Many of the California Location Installed capacity (MW)
wind farms were originally constructed in the 1980s and
are now being re-equipped with larger modern wind Germany 5432
turbines. Denmark 2281
Table 3.3.1show the installed wind-power Spain 2099
capacity worldwide in January 2001 although it is Netherlands 444
obvious that with such a rapid growth in some countries UK 391
data of this kind become out of date very quickly. Total Europe 11831
The reasons development of wind energy in some California 1622
countries is flourishing while in others it is not fulfilling Total USA 2568
the potential that might be anticipated from a simple Total World 16461
consideration of the wind resource, are complex.
Important factors include the financial-support
mechanisms for wind-generated electricity, the process by which the local planning authorities give
permission for the construction of wind farms, and the perception of the general population particularly
with respect to visual impact. In order to overcome the concerns of the rural population over the
environmental impact of wind farms there is now increasing interest in the development of sites
offshore.

| Historical Uses of Wind Pavesi - 15 -


3.4 Darrieus Wind Turbines
Most wind turbines designed for
the production of electricity have
consisted of a two or three bladed
propeller rotating around a horizontal
axis. These blades tend to be expensive,
high technology items, and the turbine
has to be oriented into the wind, another
expensive task for the larger machines.
These problems have led many
researchers in search of simpler and less
expensive machines. The variety of such
machines seems endless. One that has
seen considerable development is the
Darrieus wind turbine. The Darrieus was
patented in the United States by G. J. M.
Darrieus in 1931 [Ramler JR, Donovan
RM (1979)]. It was reinvented by
engineers with the National Research
Council of Canada in the early 1970’s.
Sandia Laboratories built a 5 m diameter
Darrieus in 1974, and has been strongly
involved with further research on the
Darrieus turbine since that time
[Sheldahl RE, Blackwell BF (1977)].
Figure 3.4.1 shows a 17 meter
Darrieus built at Sandia. The diameter of
Figure 3.4.1 Sandia Laboratories 17-m Darrieus, rated at
the blades is the same as the height, 17
60 kW in a 12.5-m/s wind.
m. The extruded aluminium blades were
made by Alcoa (Aluminium Company of
America, Alcoa Center, Pennsylvania).
This machine is rated at 60 kW in a 12.5
m/s wind. Figure 3.4.2 show one of the
blades during fabrication. Several models
of this basic machine were built during
1980.
The Darrieus has several attractive
features. One is that the machine rotates
about a vertical axis, hence does not need
to be turned into the wind. Another is
that the blades take the shape of a
jumping rope experiencing high
centrifugal forces. This shape is called
troposkein, from the Greek for turning Figure 3.4.2 Extruded aluminium blade of 17-m Darrieus.

| Historical Uses of Wind Pavesi - 16 -


rope. Since the blade operates in almost pure tension, a relatively light, inexpensive blade is sufficient.
Another advantage is that the power train, generator, and controls are all located near ground level,
hence are easier to construct and maintain. The efficiency is nearly as good as that of the horizontal
axis propeller turbine, so the Darrieus holds considerable promise as a cost effective turbine.
One disadvantage of the Darrieus is that it is not normally self starting. That is, if the turbine has
stopped during a period of low wind speeds, it will not usually start when the wind speed increases.
Starting is usually accomplished by an induction motor connected to the local utility network. This is
not necessarily a major disadvantage because the same induction motor can be used as an induction
generator to supply power to the utility network when the turbine is at operating speed. Induction
machines are simple, rugged, and inexpensive, requiring essentially no controls other than a contactor
to connect the machine to the utility network. For these reasons, they are seeing wide use as wind
turbine generators.
The first large Darrieus constructed was a 230-kW machine on Magdalen Island, Quebec, Canada
in May, 1977 by Dominion Aluminium Fabricators, Limited of Ontario, Canada. The average power
output of this machine was 100 kW over the first year of operation, which is quite good. Then a noise
was observed in the gearbox so the machine was stopped for inspection and repairs. During the
inspection process, the brakes were removed, which should have been safe because the turbine was not
supposed to be able to self start. Unfortunately, on July 6, 1978, the turbine started, and without a load
or any way of stopping it, went well over the design speed of 38 r/min. The spoilers did not activate
properly, and when the turbine reached 68 r/min a guy wire broke, letting the turbine crash to the
ground. Perhaps the main lesson learned from this accident was that the Darrieus will sometimes start
under unusual gust conditions and that braking systems need to be designed with this fact in mind.
A major design effort on Darrieus turbines has been made by Alcoa. They first designed a 5.5 m
diameter machine which would produce about 8 kW of power, but dropped that size in favour of more
economical larger machines. Other sizes developed by Alcoa include a 12.8 m diameter (30 to 60 kW),
17 m diameter (60 to 100 kW), and a 25 m diameter (300 or 500 kW depending on the gear ratio).
The Alcoa effort has been plagued by a number of accidents. A 12.8 m diameter machine
collapsed at their Pennsylvania facility on March 21, 1980, when its central torque tube started
vibrating and eventually buckled when the machine was running above rated speed. Then in April,
1981, a 25 m machine crashed in the San Gorgonio Pass east of Los Angeles [Wind Energy Report
(1981)].
The machine itself worked properly to a speed well above rated speed, but a software error in the
microcomputer controller prevented proper brake application in high winds. When the machine
rotational speed reached 60 r/min, well above the rated speed of 41 r/min, a bolt broke and allowed a
blade to flare outward and cut one of the guy wires. The machine then crashed to the ground.
Accidents like these are not uncommon in new technology areas, but they are certainly frustrating
to the people involved. It appears that the various problems are all solvable, but the string of accidents
certainly slowed the deployment of Darrieus turbines as compared with the horizontal axis turbines.
Sandia continued work on the theory of the Darrieus turbine during the 1980s, with the result that
the turbine is well understood today. It appears that there is no reason the Darrieus could not be an
important contributor to the production of power from the wind. It just needs a large aluminium
company that is willing and able to do the aluminium extrusions and possibly wait for several years
before seeing a significant return on investment.

| Historical Uses of Wind Pavesi - 17 -


3.5 Innovative Wind Turbines
Another type of turbine developed at about the same time as the Darrieus was the Savonius
turbine, developed in Finland by S. J. Savonius [Savonius, S. J. (1931)]. This is another vertical axis
machine which needs no orientation into the wind. Alternative energy enthusiasts often build this
turbine from used oil barrels by cutting the barrels in half lengthwise and welding the two halves back
together offset from one another to catch the wind. A picture of a somewhat more advanced unit
developed at Kansas State University, Manhattan, Kansas, is shown in Figure 3.5.1
The tower of the KSU Savonius was 11 m high and 6 m wide. Each rotor was 3 m high by 1.75 m
in diameter. The rotors were connected together and drove a single 5 kW, three-phase, permanent
magnet generator. At the rated wind speed of 12 m/s, the rotor speed was 103 r/min, the generator
speed was 1800 r/min, and the frequency was 60 Hz. Output voltage and frequency varied with wind
speed and load, which meant that this particular turbine could not be directly paralleled with the utility
grid. Applications for this asynchronous (not synchronized with the utility grid) electricity are limited
to electric heating and driving Wind Energy Systems by Dr. Gary L. Johnson three-phase induction
motors in situations which can tolerate variable speed operation. These include heat pumps, some water
pumps, and fans. Such applications consume large quantities of electrical energy, so variable frequency
operation is not as restrictive as it might appear. Asynchronous systems do not require complex blade
pitch, voltage, and frequency controls, hence should be less expensive. The main advantages of the
Savonius are a very high starting torque and simple construction. The disadvantages are weight of
materials and the difficulty of designing the
rotor to withstand high wind speeds. These
disadvantages could perhaps be overcome by
good engineering if the turbine efficiency were
high enough to justify the engineering effort
required.
Agreement on the efficiency of the
Savonius turbine apparently has finally been
reached a half century after its development.
Savonius claimed an efficiency of 31 per cent
in the wind tunnel and 37 per cent in free air.
However, he commented:[ Savonius, S. J.
(1931)] “The calculations of Professor Betz
gave 20 % as the highest theoretical maximum
for vertical airwheels, which under the best of
circumstances could not produce more than 10
% in practical output.” The theoretical and
experimental results failed to agree.
Unfortunately, Savonius did not specify the
shape and size of his turbine well enough for
others to try to duplicate his results.
A small unit of approximately 2 m high
by 1 m diameter was built and tested at Kansas
State University during the period 1932-1938
Figure 3.5.1 Kansas State University Savonius, [Kloeffler, R. G. and E. L. Sitz (1946)]. This
rated at 5 kW in a 12-m/s wind.

| Historical Uses of Wind Pavesi - 18 -


unit was destroyed by a high wind, but efficiencies of 35 to 40 % were claimed by the researchers.
Wind tunnel tests were performed by Sandia on 1.5 m high by 1 m diameter Savonius turbines, with a
maximum efficiency measured of 25 % for semicircular blades [Blackwell, B. F., et al. (1977)].
Different blade shapes which were tested at the University of Illinois showed a maximum efficiency of
about 35 % [Khan, M. H. (1978)]. More Savonius turbines were tested at Kansas State University, with
efficiencies reported of about 25 % [Turnquist, R. O. and F. C. Appl (1976), Johnson, G. L. (1979)]. It
thus appears that the Savonius, if properly designed, has an efficiency nearly as good as the horizontal
axis propeller turbine or the Darrieus turbine. The Savonius turbine therefore holds promise in
applications where low to medium technology is required or where the high starting torque is
important.
A chart of efficiency of five different turbine types is shown in Figure 3.5.2. The efficiency or
power coefficient varies with the ratio of blade tip speed to wind speed, with the peak value being the
number quoted for a comparison of turbines. This will be discussed in more detail in the next chapters.
It may be noticed that the peak efficiencies of the two bladed propeller, the Darrieus, and the Savonius
are all above 30 %, while the American Multiblade and the Dutch windmills peak at about 15%. These
efficiencies indicate that the American Multiblade is not competitive for generating electricity, even
though it is almost ideally suited and very competitive for pumping water.
The efficiency curves for the Savonius and the American Multiblade have been known for a long
time [Kloeffler, R. G. (1946), Savonius, S. J. (1931)]. Unfortunately, the labels on the two curves were
accidentally interchanged in some key publication in recent years, with the result that many authors
have used an erroneous set of curves in their writing. This historical accident will probably take years
to correct.
Another vertical axis machine which has interested people for many years is the Madaras rotor.
This system was invented by Julius D. Madaras, who conducted considerable tests on his idea between
1929 and 1934. This concept uses the Magnus effect, which refers to the force produced on a spinning
cylinder or sphere in a stream of air. The most familiar example of this effect is the curve ball thrown
by a baseball pitcher. The
Madaras rotor is a large
cylinder which is spun in the
wind by an electric motor.
When the wind is from the left
and the cylinder is spinning
counter clockwise as shown
Figure 3.5.3, the cylinder will
experience a lift force in the
direction shown. There will
also be a drag force in the
direction of the wind flow.
If the cylinder is
mounted on a special type of
railroad car and if the wind
speed component
perpendicular to the railroad
tracks is sufficiently strong,
Figure 3.5.2 Typical performances of wind machines.
the lift force will be adequate

| Historical Uses of Wind Pavesi - 19 -


to move the car along the tracks. The basic idea is
shown in Figure 3.5.4. The railroad car or tracked
carriage must be heavy enough that it will not
overturn due to the drag forces. Power can be
extracted from the system by electrical generators
connected to the wheels of the tracked carriage.
The cars roll around a circular or racetrack shaped
track. Twice during each orbit of a rotor car
around the track (when the wind is parallel to the
track), each spinning rotor in turn must be de-
spun to a stop, and then spun-up in the opposite
direction. This cycle is necessary in order to
assure that the propulsive force changes direction
so that all rotors are propelling the train in the
same angular direction.
Figure 3.5.3 Magnus force on a spinning The original system proposed by Madaras
cylinder. consisted of 27 m high by 6.8 m diameter
cylinders which were vertically mounted on flat
cars and rotated by electric motors to convert wind energy to Magnus-effect forces. The forces would
propel an endless train of 18 cars around a 460 m diameter closed track. Generators geared to the car
axles were calculated to produce up to 18 MW of electric power at a track speed of 8.9 m/s in a wind
speed of 13 m/s.
More recent studies [Whitford, D. H., J. et al. (1978), Whitford, D. H. and J. E. Minardi (1979)]
have shown that energy production is greater with a racetrack shaped plant perhaps 3 km wide by 18
km long which is oriented perpendicular to the
prevailing winds. This modern design includes
cylinders 4.9 m in diameter by 38.1 m tall, cars
with a length of 19.2 m and a width of 17.4 m,
and a track with 11 m between rails. Individual
cars would have a mass of 328,000 kg. Each
rotor would be spun with a 450 kW, 500 volt
dc motor. Each of the four wheels would drive
a 250 kW induction generator. There would be
about 200 cars on the track with a total rating
of about 200 MW. Power would be extracted
from the system by a 4160 V, three-phase, 500
A overhead trolley bus.
Cost estimates for the electricity costs
from this large system were comparable to
those from the MOD-1. Wind tunnel tests and
Figure 3.5.4 Madaras concept for generating field tests on a rotating cylinder on a fixed
electricity. platform indicate that the concept will work.
The questions remain whether the aerodynamic, mechanical, and electrical losses will be acceptable
and whether the reliability will be adequate.
Only a major development effort can answer these questions and there will probably not be

| Historical Uses of Wind Pavesi - 20 -


sufficient interest in such a development if
the horizontal axis wind turbines meet the
basic requirements for cost and reliability.
All the wind turbines discussed thus
far have a problem with capital costs.
Although these machines work
satisfactorily, capital costs are large. The
Darrieus may become more cost effective
than the two-bladed propeller turbine, but
neither is likely to produce really
inexpensive electricity. There is a desire
for a breakthrough, whereby some new
and different concept would result in
substantial cost reductions. One candidate
for such a wind machine is the augmented
vortex turbine studied by James Yen at
Grumman Aerospace Corporation [Yen, J.
T. (1977)].
An artist’s concept of the machine is
shown in Figure 3.5.5. The turbine tower
has vertical vanes which direct the wind
into a circular path around the inside of the
tower. Wind blowing across the top of the
tower tends to pull the air inside in an
upward direction, causing the entering air
Figure 3.5.5 Augmented vortex turbine. to flow in a spiral path. This spiral is a
vortex, which is characterized by a high
speed, low pressure core. The vortex is basically that of a confined tornado. The pressure difference
between the vortex core and outside ambient air is then used to drive a relatively small, high speed
turbine at the base of the tower. The vortex machine is extracting power from pressure differences or
the potential energy in the air, rather than directly from the kinetic energy of the moving air. The
potential energy in the air due to pressure is vastly more than the kinetic energy of the air in moderate
wind speeds, so there is a possibility of large energy outputs for a given tower size which could result
in very inexpensive electricity.
One problem with the vortex machine is the potential for spawning tornadoes. If the vortex
extending out of the top of the tower should become separated from the tower, grow a tail, and become
an actual tornado, a permanent shutdown would be highly probable. In fact, based on the experience of
the nuclear industry, fear of such an occurrence may prevent the implementation of such a wind
machine.
Many other wind machines have been invented over the last few hundred years. The propeller
type and the Darrieus have emerged as reasonably reliable, cost competitive machines which can
provide a significant amount of electrical energy. Barring a major breakthrough with another type of
wind machine, we can expect to see a wide deployment of these machines over the next few decades.

| Historical Uses of Wind Pavesi - 21 -


3.6 Current status and future prospects
Owing to our commitment to reduce GHG emissions and provide adequate energy to the
developing world, efforts are being made to supplement our energy base with renewable sources.
Several countries have already formulated policy frameworks to ensure that renewables play an
impressive role in the future energy scenario.
For example, the European Union targets to meet 22 per cent of their demand from renewables by
2010. Wind, being the commercially viable and economically competitive renewable source, is going
to be the major player in
meeting this target.
Wind is the world’s
fastest growing energy
source today and it has
been retaining this position
consecutively for the last
five years. The global wind
power capacity has
increased by a factor of 4.2
during the last five years.
The total global installed
capacity is 282587 MW in
2012. Installed capacity in
different regions is shown
in Figure 3.6.1. Over 73
per cent of the global
installations are in Europe.
Figure 3.6.1 Installed Wind energy capacity (MW) in different regions Germany is the European
(2012) leader, followed by Spain,
United Kingdom and Italy.
The five countries, leading in wind energy generation are listed in Table 3.6.1 and Table 3.6.3

Table 3.6.1 Global leaders in wind energy generation 2012


Country Installed capacity, MW
China 75324
United States of America 60007
Germany 31308
Spain 22796
India 18421
United Kingdom 8445
India 8144
With the increasing thrust on renewables and reducing cost of wind generated electricity, the
growth of wind energy will continue in the years to come. According to European Wind Energy
Association (EWEA), wind with its expected 230,000 MW installation, can supply 12 per cent of the
global energy demand by 2010 [de Azua CR, Colasimone L (2003)]. This indicates a market worth

| Historical Uses of Wind Pavesi - 22 -


around 25 billion Euros. The installed capacity Table 3.6.2 Wind energy generation
may reach a level of 1.2 million MW by 2020. Installed capacity MW 2015 2005
In tune with the growth of the industry, the Valle D’Aosta 0 0
wind energy technology is also changing. One Piemonte 12,5 0
apparent change is the shift towards offshore Lombardia 0 0
installations. Several ambitious offshore Trentino Alto Adige 2,5 1,2
projects are in the pipeline. For example, 20 Veneto 1,35 0
offshore projects are planned to be installed in Friuli Venezia Giulia 0 0
UK by 2006, with a total capacity of 1400 Liguria 22,6 4,8
MW [Zaaijer M, Henderson A (2003)]. Emilia Romagna 16,3 3,5
In Germany, around 30 offshore projects
Toscana 45 1,8
worth 60,000 MW are in various stages of
Marche 0 0
processing. In United States also, the offshore
Umbria 1,5 1,5
activities are intensifying. Another trend in the
industry is to go for larger machines. As Lazio 9 9
bigger turbines are cheaper on a unit kW basis, Abruzzo 225 158,2
the industry is growing from MW to multi- Molise 372 35,3
MW scale. The 2 MW+ sector is rapidly Campania 814 311,5
growing. Several manufactures like the RE Puglia 1.286 284,9
Power Systems AG are coming up with Basilicata 279 85,5
turbines of even 5 MW size. The RE Power Calabria 589 0
model is equipped with a huge 125 m rotor Sardegna 673 241,6
having each blade weighing around 19 tonnes Sicilia 1.449 230,4
[de Vries E (2003)]. Efforts are also on to Italy 5.797 1.369,4
reduce the total head mass (THM) which is the
total mass of nacelle and rotor. Reduction in THM has positive impact on system dynamics. By clever
engineering design, NEG Micon could restrict the THM of their 4.2 MW model to 214 tonnes, which is
a remarkable achievement.
Owing to the active grid support and better efficiency, the variable speed option with double fed
induction generator is getting more prominence in the industry.
Another innovative concept that may prove effective in future is the direct drive machines.
Over the last 10 years there has been a continuous increase in the rotor diameter of commercially
available wind turbines from around 30 m to more than 60 m. A doubling of the rotor diameter leads to
a four-times increase in power output. The influence of the wind speed is, of course, more pronounced
with a doubling of wind speed leading to an eight-fold increase in power. Thus there have been
considerable efforts to ensure that wind farms are developed in areas of the highest wind speeds and the
turbines optimally located within wind farms. In certain countries very high towers are being used
(more than 60–80 m) to take advantage of the increase of wind speed with height.
In the past a number of studies were undertaken to determine the ‘optimum’ size of a wind
turbine by balancing the complete costs of manufacture, installation and operation of various sizes of
wind turbines against the revenue generated (Molly et al., 1993). The results indicated a minimum cost
of energy would be obtained with wind turbine diameters in the range of 35–60 m, depending on the
assumptions made. However, these estimates would now appear to be rather low and there is no
obvious point at which rotor diameters, and hence output power, will be limited particularly for
offshore wind turbines.

| Historical Uses of Wind Pavesi - 23 -


The Italy is still now on the fringe of the wind energetic strategy Table 3.6.3.

Table 3.6.3 EU leaders in wind energy generation


Country Installed capacity MW
December 2010 December 2011 December 2012
Germaniy 27.214 29,071 31,308
Spain 20.676 21,674 22,796
UK 5.204 6,556 8,445
Italy 5.797 6,878 8,144
France 5.660 6,807 7,564

| Historical Uses of Wind Pavesi - 24 -


3.7 Reference

Blackwell, B. F., R. E. Sheldahl, and L. V. Feltz: Wind Tunnel Performance Data for Two- and Three-
Bucket Savonius Rotors, Sandia Laboratories Report SAND 76-0131, July 1977.
CEU, (1997). ‘Energy for the future, renewable sources of energy – White Paper for a Community
Strategy and Action Plan’. COM (97) 559 final.
de Azua CR, Colasimone L (2003) Record growth for global wind power in 2002; 28% increase, wind
technology worth $7.3 billion installed last year. AWEA-EWEA News release, Global Wind
Power Installations, http://www.ewea.org
de Vries E (2003) Wind turbine technology trends – review 2003. Renewable Energy World 6(4): 154-
167
Freris, L. L. (ed.), (1990). Wind energy conversion systems. Prentice Hall, New York, US.
Golding, E. W. (1955). The generation of electricity from wind power. E. & F. N. Spon (reprinted R. I.
Harris, 1976).
Golding, E. W. (1976) The generation of electricity by wind power. Halsted Press, New York.
Johnson, G. L.: “Preliminary Results of a 5-kW Savonius Wind Turbine Test,” USDADOE Workshop
on Wind Energy Application in Agriculture, Ames, Iowa, May 15-17, 1979.
Johnson G. L. (2001) Wind energy systems. http://www.rpc.com.au
Khan, M. H.: “Model and Prototype Performance Characteristics of Savonius Rotor Windmill,” Wind
Engineering, Vol. 2, No. 2, 1978, pp. 75-85.
Kloeffler, R. G. Sitz E. L.(1946): Electric Energy from Winds, Kansas State College of Engineering
Experiment Station Bulletin 52, Manhattan, Kans., September 1, 1946.
Molly, J. P. Keuper, A. and Veltrup, M., (1993). ‘Statistical WEC design and cost trends’. Proceedings
of the European Wind Energy Conference, pp 57–59.
Putnam, P. C. (1948) Power from the wind. Van Nostrand, New York.
Ramler JR, Donovan RM (1979) Wind turbines for electric utilities: Development status and
economics. DOE/NASA/1028-79/23, NASA TM-79170, AIAA-79-0965
Savonius, S. J. (1931): “The S-Rotor and Its Applications,” Mechanical Engineering, Vol. 53, No. 5,
May 1931, pp. 333-338.
Sheldahl RE, Blackwell BF (1977) Free-air performance tests of a 5-meter-diameter darrieus turbine.
Sandia Laboratories Report SAND 77-1063
Sorensen B (1995) History of, and recent progress in, wind-energy utilization. Annual Review of
Energy and the Environment 20(1): pp. 387-424.
Spera, D. A. (1994) Wind-turbine technology, fundamental concepts of wind-turbine engineering.
ASME Press, New York, US.
The Windicator (2005) Wind energy facts and figures from wind power monthly. Windpower Monthly
News Magazine, Denmark, USA: 1-2
Turnquist, R. O. and F. C. Appl: “Design and Testing of a Prototype Savonius Wind Machine,”
Frontiers of Power Technology Conference, Oklahoma State University, Stillwater, Okla.,
October 27-28, 1976.
Vargo, D. J. (1974): Wind Energy Development in the 20th Century, NASA Technical Memorandum
NASA TM X-71634, September, 1974.
Third Wind Energy Workshop (1977), Washington, D. C., September, 1977, CONF 770921 D.C.
Zervos, A. (2000) ‘European targets, time to be more ambitious?’ Windirections, 18–19.

| Historical Uses of Wind Pavesi - 25 -


Zaaijer M, Henderson A (2003) Offshore update – A global look at offshore wind energy. Renewable
Energy World 6(4): 102-119
Whitford, D. H., J. E. Minardi, B. S. West, and R. J. Dominic: An Analysis of the Madaras Rotor
Power Plant: An Alternative Method for Extracting Large Amounts of Power from the Wind,
DOE Report DSE-2554-78/2, Vol. 2, June 1978.
Whitford, D. H. and J. E. Minardi: “Utility-Sized Wind-Powered Electric Plants Based on the Madaras
Rotor Concept,” Wind Energy Innovative Systems Conference Proceedings, SERI/TP-245-184,
May 23-25, 1979, pp. 71-81.
Wind Energy Report, April 1981.
Yen, J. T.: “Summary of Recent Progress on Tornado-Type Wind Energy System,” Third Wind Energy
Workshop Proceedings, Washington, D.C., September 1977, CONF 770921, pp. 808-818.

| Historical Uses of Wind Pavesi - 26 -


4 Wind Characteristics
The earth’s atmosphere can be modelled as a gigantic heat engine. It extracts energy from one
reservoir (the sun) and delivers heat to another reservoir at a lower temperature (space).
In the process, work is done on the gases in the atmosphere and upon the earth-atmosphere
boundary. There will be regions where the air pressure is temporarily higher or lower than average.
This difference in air pressure causes atmospheric gases or wind to flow from the region of higher
pressure to that of lower pressure. These regions are typically hundreds of kilometres in diameter.
Solar radiation, evaporation of water, cloud cover, and surface roughness all play important roles
in determining the conditions of the atmosphere. The study of the interactions between these effects is a
complex subject called meteorology, which is covered by many excellent textbooks.[Cole, F. W.
(1970), Donn, W. L. (1965), Riehl, H. (1965)] Therefore only a brief introduction to that part of
meteorology concerning the flow of wind will be given in this text.

4.1 Meteorology of Wind


The basic driving force of air movement is a difference in air pressure between two regions. This
air pressure is described by several physical laws. One of these is Boyle’s law, which states that the
product of pressure and volume of a gas at a constant temperature must be a constant, or
p1V1=p2V2 (5.1.1)
Another law is Charles’ law, which states that, for constant pressure, the volume of a gas varies
directly with absolute temperature.
V1 V2
 (5.1.2)
T1 T2
If a graph of volume versus temperature is made from measurements, it will be noticed that a zero
volume state is predicted at −273.15oC or 0 K.
The laws of Charles and Boyle can be combined into the ideal gas law
pV = nRT (5.1.3)
In this equation, R is the universal gas constant, T is the temperature in Kelvin, V is the volume
of gas in m3, n is the number of kilomoles of gas, and p is the pressure in Pascal (N/m2). At standard
conditions, 0oC and one atmosphere, one kilomole of gas occupies 22.414 m3 and the universal gas
constant is 8314.5 J/(kmol K) where J represents a joule or a Newton meter of energy. The pressure of
one atmosphere at 0°C is then
(8314.5J/(kmol K))(273.15 K)
 101, 325 Pa (5.1.4)
22.414 m3
One kilomole is the amount of substance containing the same number of molecules as there are
atoms in 12 kg of the pure carbon nuclide 12C. In dry air, 78.09 % of the molecules are nitrogen, 20.95
% are oxygen, 0.93 % are argon, and the other 0.03 % are a mixture of CO2, Ne, Kr, Xe, He, and H2.
This composition gives an average molecular mass of 28.97, so the mass of one kilomole of dry air is
28.97 kg. For all ordinary purposes, dry air behaves like an ideal gas.
The density ρ of a gas is the mass m of one kilomole divided by the volume V of that kilomole.

| Wind Characteristics Pavesi - 27 -



m (5.1.5)
V
The volume of one kilomole varies with pressure and temperature as specified by Eq. (5.1.3)
When we insert Eq. (5.1.3) into Eq. (5.1.5), the density is given by
mp 3.484p
  kg/m3 (5.1.6)
RT T
where p is in kPa and T is in Kelvin. This expression yields a density for dry air at standard
conditions of 1.293 kg/m3.
The common unit of pressure used in the past for meteorological work has been the bar (100 kPa)
and the millibar (100 Pa). In this notation a standard atmosphere was referred to as 1.01325 bar or
1013.25 millibar. Atmospheric pressure has also been given by the height of mercury in an evacuated
tube. This height is 760 millimetres of mercury for a standard atmosphere. It may be worth noting here
that several definitions of standard conditions are in use. The chemist uses 0°C as standard temperature
while engineers have often used 20°C or 25°C as standard temperature. We shall not debate the
respective merits of the various choices, but note that some physical constants depend on the definition
chosen, so that one must exercise care in looking for numbers in published tables. In this text, standard
conditions will always be 0°C and 101.3 kPa.
Within the atmosphere, there will be large regions of alternately high and low pressure. These
regions are formed by complex mechanisms, which are still not fully understood. Solar radiation,
surface cooling, humidity, and the rotation of the earth all play important roles. From the point of view
of wind energy, the most striking characteristic of the wind resource is its variability. The wind is
highly variable, both geographically and temporally. Furthermore this variability persists over a very
wide range of scales, both in space and time. The importance of this is amplified by the cubic
relationship to available energy.
On a large scale, spatial variability describes the fact that there are many different climatic
regions in the world, some much windier than others. These regions are largely dictated by the latitude,
which affects the amount of insolation. Within any one climatic region, there is a great deal of variation
on a smaller scale, largely dictated by physical geography – the proportion of land and sea, the size of
land masses, and the presence of mountains or plains for example. The type of vegetation may also
have a significant influence through its effects on the absorption or reflection of solar radiation,
affecting surface temperatures, and on humidity.
More locally, the topography has a major effect on the wind climate. More wind is experienced
on the tops of hills and mountains than in the lee of high ground or in sheltered valleys, for instance.
More locally still, wind velocities are significantly reduced by obstacles such as trees or buildings.
At a given location, temporal variability on a large scale means that the amount of wind may vary
from one year to the next, with even larger scale variations over periods of decades or more. These
long-term variations are not well understood, and may make it difficult to make accurate predictions of
the economic viability of particular wind-farm projects, for instance.
On time-scales shorter than a year, seasonal variations are much more predictable, although there
are large variations on shorter time-scales still, which although reasonably well understood, are often
not very predictable more than a few days ahead. These ‘synoptic’ variations are associated with the
passage of weather systems. Depending on location, there may also be considerable variations with the
time of day (diurnal variations) which again are usually fairly predictable. On these time-scales, the
predictability of the wind is important for integrating large amounts of wind power into the electricity

| Wind Characteristics Pavesi - 28 -


Figure 4.1.1 Wind Spectrum Farm Brookhaven Based on Work by van der Hoven (1957)
network, to allow the other generating plant supplying the network to be organized appropriately.
On still shorter time-scales of minutes down to seconds or less, wind-speed variations known as
turbulence can have a very significant effect on the design and performance of the individual wind
turbines, as well as on the quality of power delivered to the network and its effect on consumers.
Van der Hoven (1957) constructed a wind-speed spectrum from long- and short-term records at
Brookhaven, New York, showing clear peaks corresponding to the synoptic, diurnal and turbulent
effects referred to above (Figure 4.1.1). Of particular interest is the so-called ‘spectral gap’ occurring
between the diurnal and turbulent peaks, showing that the synoptic and diurnal variations can be treated
as quite distinct from the higher-frequency fluctuations of turbulence. There is very little energy in the
spectrum in the region between 2 h and 10 min.
The flux of solar radiation incident on a surface placed at the top of the atmosphere depends on
time (t) and geographical location (latitude φ and longitude λ), and on the orientation of the surface,

0   t, ,    S  t  cos   t, ,  
ESW (5.1.7)
Here S(t) is the “solar constant” at the distance of the Earth (being a function of time due to
changes in the Sun–Earth distance and due to changes in solar luminosity), and θ is the angle between
the incident solar flux and the normal to the surface considered. The subscript “0” on the short-
wavelength flux ESW through the surface indicates that the surface is situated at the top of the
atmosphere, and "+” indicates that only the positive flux (in the "inward” direction) is considered. For a
horizontal surface, θ is the zenith angle z, obtained by
cosz = sinδ sinφ + cosδ cosφ cosω (5.1.8)
where δ is the declination of the Sun and ω the hour angle of the Sun (see Figure 4.1.3). The

| Wind Characteristics Pavesi - 29 -


declination is given approximately by (Cooper, 1969)

Figure 4.1.2 Schematic energy cycle without anthropogenic interference. The energy flows are in
TW (1012W). Numbers in parentheses are uncertain or rounded off (with use of Odum, 1972; Budyko,
1974; Gregg, 1973; Lorenz, 1967; Manabe, 1969). The earth receives around 1.7 1014 kW of power
from the sun in the form of solar radiation. This radiation heats up the atmospheric air.

| Wind Characteristics Pavesi - 30 -


Figure 4.1.3 Coordinates in a Figure 4.1.4 Daily average radiation,
geocentric picture. P represents the E SW
0 (Wm-2), incident at the top of the
North Pole, Z zenith and φ the Earth’s atmosphere, as a function of time
geographical latitude. The unit radius of the year and latitude (independent of
great circle shown through P and Z is longitude). Latitude is taken as positive
the median of the observer. S towards north (in this and the following
represents the direction to the Sun, figures) (based on NCEP-NCAR, 1998).
and the solar coordinates are the
declination δ and the hour angle ω.
Also shown is the zenith angle z
(equal to π/2 minus the height  284  day 
 of0.4093sin
the  2 365  (5.1.9)
Sun over the horizon).  
where day is the actual day’s number in the year. The hour angle (taken as positive in the
mornings) is related to the local time tzone (in hours) by
ω = 2π (12 − tzone)/24 − (λ - λzone) − TEQ (5.1.10)
Here λzone is the longitude of the meridian defining the local time zone (longitudes are taken
positive towards east, latitudes and declinations positive towards north, all angles are in radians). The
correction TEQ (“equation of time”) accounts for the variations in solar time caused by changes in the
rotational and orbital motion of the Earth (e.g. due to speed variations in the elliptical orbit and to the
finite angle (obliquity) between the axis of rotation and the normal to the plane or orbital motion). The
main part of TEQ remains unaltered at the same time in subsequent years. It may then be expressed as a
function of the day of the year (see Duffie and Beckman, 1974),
TEQ = – 0.0113 – 0.0019 × day, if day ≤ 20,
= – 0.0227 – 0.0393 cos(0.0357(day−43)), if 20 < day ≤ 135,
= – 0.0061 + 0.0218 cos(0.0449(day−135)), if 135 < day ≤ 240,
= 0.0275 + 0.0436 cos(0.0360(day−306)), if 240 < day ≤ 335,
= – 0.0020 × (day−359), if day > 335.
The resulting daily average flux on a horizontal surface,
1
0 
ESW  0  t, ,   dt
ESW (5.1.11)
24 day

| Wind Characteristics Pavesi - 31 -


which is independent of λ, is shown in Figure 4.1.4. The maximum daily flux is about 40% of the flux
on a surface fixed perpendicular to the direction of the Sun. It is obtained near the poles during
midsummer, with light close to 24 hours of day but a zenith angle of π/2 – max(|δ|) ≈ 67° (cf.Figure
4.1.3, |δ| being the absolute value of δ). Daily fluxes near to 30% of the solar constant are found
throughout the year at the Equator. Here the minimum zenith angle comes close to zero, but day and
night are about equally long.
A fraction of the incoming solar radiation is reflected back into space. This fraction is called the
albedo a0 of the Earth–atmosphere system. The year and latitude–longitude average of a0 is about 0.35.
This is composed of about 0.2 from reflection on clouds, 0.1 from reflection on cloudless atmosphere
(particles, gases) and 0.05 from reflection on the Earth’s surface. Figure 4.1.5 gives the a0 distribution
for two months of 1997, derived from ongoing satellite measurements (cf. Wilson and Matthews, 1971;
Raschke et al., 1973, NCEPNCAR, 1998).
The radiation absorbed by the Earth–atmosphere system is

0  1  a 0 
A 0  Esw (5.1.12)
Since no gross change in the Earth’s temperature is taking place from year to year, it is expected
that the Earth–atmosphere system is in radiation equilibrium, i.e. that it can be ascribed an effective
temperature T0, such that the black-body radiation at this temperature is equal to –A0,

Figure 4.1.5 a, b Top of the atmosphere Figure 4.1.6 Latitude dependence (averaged over
albedo for the Earth atmosphere system, longitudes and over the year) of incident ( E sw 0  ) and
derived from satellite data for January (top)
outgoing long-wavelength ( Esw
0  ) radiation at the
and July (bottom) 1997 (NCEP-NCAR, 1998).
top of the atmosphere, the amount of radiation
absorbed (A0) and reflected (R0) by the Earth−
atmosphere system, and the net radiation flux E0 at
the top of the atmosphere (in Wm-2; based on
Sellers, 1965; Budyko, 1974).

| Wind Characteristics Pavesi - 32 -


1
1  4

 4 S 1  a o  
T0     253K (5.1.13)
  
 
Here S is the solar constant and σ = 5.7×10-8 Wm-2 K-4 is Stefan’s constant. The factor 1/4 comes
because the incoming radiation on the side of the Earth facing the Sun is the integral of S cosθ over the
hemisphere, while the outgoing radiation is uniform over the entire sphere (by definition when it is
used to define an average effective temperature). The net flux of radiation at the top of the atmosphere
can be written:

0  1  a 0   E 0
E 0  ESW IW
(5.1.14)

where E 0hw  E lw
0  on average equals T0 (recall that fluxes away from the Earth are taken as
4

negative) and E0 on average (over the year and the geographical position) equals zero.
The fact that T0 is 34 K lower than the actual average temperature at the Earth’s surface indicates
that heat is accumulated near the Earth’s surface, due to re-radiation from clouds and atmosphere. This
is called the greenhouse effect and it will be further illustrated below, when discussing the radiation
fluxes at the ground.
The net radiation flux at the top of the atmosphere, E0, is a function of time as well as of
geographical position. The longitudinal variations are small. Figure 4.1.7 gives the latitude distribution
of E0, A0 and R0 , where the reflected flux is
𝑆𝑊
𝑅0 = 𝐸0+ 𝑎0 (5.1.15)
as well as E lw
o .
Although R0 is fairly independent of φ, the albedo a0 increases when going towards the poles.
Most radiation is absorbed at low latitudes, and the net flux exhibits a surplus at latitudes below 40°
and a deficit at higher absolute latitudes. Since no progressive increase in equatorial average
temperatures, nor any similar decrease in polar temperatures, is taking place, there must be a transport
of energy from low latitudes in the directions of both poles. The mechanisms for producing this flow
are ocean currents and winds. The necessary amount of poleward energy transport, which is
immediately given by the latitude variation of the net radiation flux (since radiation is the only mode of
energy transfer at the top of the atmosphere), is given in Figure 4.1.7, which also indicates the
measured share of ocean transport.
The total amount of energy transported is largest around |φ | = 40°, but since the total
circumference of the latitude circles diminishes as latitude increases, the maximum energy transport
across a unit length of latitude circle is found at |φ | ≈ 50°.
The net radiation fluxes discussed showed the existence of energy transport processes other than
radiation. Energy in the form of heat is transferred from the land or ocean surface by evaporation or
conduction, and from atmosphere to surface by precipitation, by friction and again by conduction in
small amounts. These processes exchange sensible and latent heat between the atmosphere and the
oceans and continents. The exchange processes in the atmosphere include condensation, evaporation
and a small amount of conduction. In addition, energy is added or removed by transport processes, such
as convection and advection. The turbulent motion of the convection processes is often described in

| Wind Characteristics Pavesi - 33 -


terms of overlaying eddies of
various characteristic sizes. The
advective motion is the result of
more or less laminar flow. All
motion in the atmosphere is
associated with friction
(viscosity), so that kinetic energy
is constantly transformed into
heat. Thus the general circulation
has to be sustained by renewed
input of energy. The same
processes are involved in the
circulation in the oceans, but the
quantitative relations between the
different processes are different,
owing to the different physical
structure of air and water (density,
viscosity, etc.). As mentioned
earlier, the source of energy for
the transport processes is the
latitude variation of the net
radiation flux. Additional Figure 4.1.7 The poleward energy transfer required by the
transport processes may take latitude variation of the net energy flux shown in Figure 4.1.6.
place on the continents, including The dashed curve represents the same quantity divided by the
river and surface run-off as well length of the latitude circle, i.e. the energy transport required
as ground water flow. Direct heat across a unit length of the latitude circle. These figures are annual
transport in dry soil is very small, and longitudinal averages (based on Sellers, 1965). The ocean
because of the smallness of the transport contribution is indicated as a dark blurred region, based
heat conductivity. on measurements with estimated uncertainties (Trenberth and
The amount of energy stored Solomon, 1994).

Figure 4.1.8 Altitude dependence of temperature, pressure and density for standard atmosphere.

| Wind Characteristics Pavesi - 34 -


in a given volume of air, water or soil may be written
Wstored  W pot  W kin  Wsens  W lat (5.1.16)
where Wpot is the geopotential energy, Wkin is the kinetic energy of external motion (flow), Wsens is
the amount of sensible heat stored (internal kinetic motion), and Wlat is the amount of latent heat, such
as the energies involved in phase changes (solid to quid, fluid to gaseous, or other chemical
rearrangement).
The zero point of stored energy is of course arbitrary. It may be taken as the energy of a state
where all atoms have been separated and moved to infinity, but in practice it is customary to use some
convenient average state to define the zero of energy.
The geopotential energy may be written
W pot    r  g  r  r (5.1.17)

where  is the density (of air, water or soil) and g the gravitational constant at the distance r from the
Earth’s centre (eventually r is replaced by z = r – rs measured from the surface rs). Both  and g further
have a weak dependence on the geographical location (, ). The kinetic energy Wkin receives
contributions both from the mean flow velocity and from turbulent (eddy) velocities.
The sensible heat is
W sens  cp T (5.1.18)
where the heat capacity cP (at constant pressure) depends on the composition, which again may be a
function of height and position. All quantities may also depend on time. The latent heat of a given
constituent such as water may be written
W lat  Lm  m w  m v   L v m (5.1.19)
where Lv (2.27 J kg-1 for water) is the latent heat of vaporisation and Lm (0.33 J kg-1 for ice) is the
latent heat of melting mv is the mixing ratio (fractional content by volume) of water vapour and mw the
mixing ratio of liquid water (applicable in oceans and soil as well as in the atmosphere). For other
constituents it may be necessary to specify the chemical compound in which they enter in order to
express the content of latent energy.
Typical variations of density, pressure and temperature, as functions of height, are shown in
Figure 4.1.8. Common names for the different layers are indicated. They are generally defined by
turning points in the temperature profile (“pauses”), or by the lower border of constant temperature
regions if no sharp change in the sign of the temperature gradient takes place. Actual temperature
profiles are not independent of latitude and season, and do not exhibit sharp “turning points”, as
indicated in Figure 4.1.9.
The intensity of this heating will be more at the equator (0° latitude) as the sun is directly
overhead. Air around the poles gets less warm, as the angle at which the radiation reaches the surface is
more acute. The density of air decreases with increase in temperature. Thus, lighter air from the equator
rises up into the atmosphere to a certain altitude and then spreads around. This causes a pressure drop
around this region, which attracts the cooler air from the poles to the equator. The wind described
above, which is driven by the temperature difference, is called the geostrophic wind, or more
commonly the global wind Figure 4.1.10. Global winds, which are not affected by the earth surface, are
found at higher altitudes.

| Wind Characteristics Pavesi - 35 -


In order for a high pressure region to be maintained while air is leaving it at ground level, there
must be air entering the region at the same time. The
only source for this air is above the high pressure
region. That is, air will flow down inside a high
pressure region (and up inside a low pressure region)
to maintain the pressure. This descending air will be
warmed adiabatically (i.e. without heat or mass
transfer to its surroundings) and will tend to become
dry and clear. Inside the low pressure region, the
rising air is cooled adiabatically, which may result in
clouds and precipitation. This is why high pressure
regions are usually associated with good weather and
low pressure regions with bad weather.
The external motion which takes place in the
atmosphere (as opposed to the internal motion of
molecules, defining the temperature) calls for
mechanisms by which kinetic energy is created.
Friction is an irreversible process by which kinetic
energy is transformed into sensible heat,
accompanied by an increase in entropy. External
forces, such as the Coriolis force impressed by the
rotation of the Earth, do not create or destroy kinetic
energy. Neither do the processes by which latent heat
is gained or lost. Thus the only sources left to
Figure 4.1.9 a,b Average temperature in accomplish the maintenance of external kinetic
the atmosphere as a function of height and motion despite the frictional damping are the
latitude, for January (above) and July (below) potential energy and the energy of sensible heat,
1995, both at longitude zero (based on NCEP- which can both be transformed into kinetic energy by
NCAR, 1998). adiabatic and reversible processes. The familiar
examples of these transformations are gravitationally
falling matter and buoyantly rising hot matter (or
descending cool matter). In order to estimate the rate
of creation and destruction of kinetic energy in the
atmosphere, it is convenient first to collect the
equations governing the general motion of the
atmosphere.
These are the equation of motion,
dv
  grad p  Fviscous  Fext (5.1.20)
dt
where the frictional force Fviscous depends on the
internal properties of the fluid (the air), and the
external force Fext comprises the Coriolis force and
the gravitational force, provided a coordinate system
Figure 4.1.10 Geotrophic wind is employed which follows the rotation of the Earth;

| Wind Characteristics Pavesi - 36 -


the equation of continuity,

div  v    (5.1.21)
t
and the law of energy conservation (first law of thermodynamics),
dW total
 Work  Head during timedt (5.1.22)
dt
The work performed on the system is that of the forces appearing on the right-hand side of the
equation of motion, and the heat added is partly due to radiation, partly to molecular conduction
processes. Wstored contains kinetic, potential and sensible heat energy. The latent heat may be omitted in
Wstored, provided any condensation or other process releasing latent heat is included in the heat added
on the right-hand side, at the time when the latent heat is transformed into sensible heat.
The above equations would have to be complemented by the equation of state (e.g. the ideal gas
law) and the source function specifying the external heat sources. These quantities are not independent
of the constitution of the air, and in general the above equations are coupled to a large number of
equations specifying the time development of the atmospheric composition (water vapour, ozone, etc.).
In practice, a simplified system of equations is used, based on an analysis of the relative
importance of different terms, in particular of the forces, and depending on the scale of motion, which
is considered interesting. The dominance of horizontal wind fields in the large-scale atmospheric
motion makes it possible to simplify the vertical (z) component of the equation of motion, neglecting
also the corresponding component of the viscous and Coriolis forces. This approximation leads to the
hydrostatic equation,
p
 g (5.1.23)
z
where g is the average acceleration of the Earth’s gravity (neglecting anisotropy in the mass
distribution inside the Earth and the latitude dependence associated with Coriolis corrections), g ≈
GM(r)/r2 ≈ 9.8 ms-2. This does not imply that terms in the equation of motion such as dw/dt, where w is
the z component of v, may be neglected in other circumstances, but only that they are small in
comparison with the two terms kept in the equation expressing the hydrostatic approximation. This
equation states that the pressure gradient at any infinitesimal horizontal area must balance the weight of
the entire air column above it, just as if there were no motion in the atmosphere.
A line drawn through points of equal pressure on a weather map is called an isobar. These
pressures are corrected to a common elevation such as sea level. The horizontal pressure difference
provides the horizontal force or pressure gradient which determines the speed and initial direction of
wind motion. In describing the direction of the wind, we always refer to the direction of origin of the
wind. That is, a north wind is blowing on us from the north and is going toward the south. The greater
the pressure gradient, the greater is the force on the air, and the higher is the wind speed. Since the
direction of the force is from higher to lower pressure, and perpendicular to the isobars, the initial
tendency of the wind is to blow parallel to the horizontal pressure gradient and perpendicular to the
isobars. However, as soon as wind motion is established, a deflective force is produced which alters the
direction of motion. This force is called the Coriolis force.

| Wind Characteristics Pavesi - 37 -


Figure 4.1.11 Coriolis force
The Coriolis force is due to the earth’s rotation under a
moving particle of air. From a fixed observation point in space
air would appear to travel in a straight line, but from our
vantage point on earth it appears to curve. To describe this
change in observed direction, an equivalent force is postulated.
The basic effect is shown in Figure 4.1.11. The two
curved lines are lines of constant latitude, with point B located
directly south of point A. A parcel of air is moving south at
point A. If we can imagine our parcel of air to have zero air
friction, then the speed of the parcel of air will remain constant
with respect to the ground. The direction will change, however,
because of the earth’s rotation under the parcel. Figure 4.1.12 Wind forces in a low-
At point A, the parcel has the same eastward speed as the pressure area
earth. Because of the assumed lack of friction, it will maintain
this same eastward speed as it moves south. The eastward
speed of the earth increases, however, as we move south (in the
Northern Hemisphere). Therefore, the parcel will appear to
have a westward component of velocity on the latitude line
passing through point B. During the time required for the
parcel to move from the first latitude line to the second, point
A has moved eastward to point A’ and point B has moved
eastward to point B’. The path of the parcel is given by the
dashed line. Instead of passing directly over point B’ which is
directly south of point A’, the parcel has been deflected to the Figure 4.1.13 Wind forces in a high-
right and crosses the second latitude line to the west of B’. The pressure area

| Wind Characteristics Pavesi - 38 -


total speed relative to the earth’s surface
remains the same, so the southward
moving component has decreased to allow
the westward moving component of speed
to increase.
It can be shown that a parcel of air
will deflect to its right in the Northern
Hemisphere, regardless of the direction of
travel. This is not an obvious truth, but the
spherical geometry necessary to prove the
statement is beyond the scope of this text.
We shall therefore accept it on faith. Under
Figure 4.1.14 Illustration of the geostrophic wind; Fp, the influence of Coriolis forces, the air
pressure force on the air; Fp, Coriolis force move almost parallel to the isobars. Thus,
in the northern hemisphere, wind tends to
rotate clockwise where as in the southern
hemisphere the motion is in the anti-
clockwise direction.
Another statement we shall accept on
faith is that the deflection of the parcel of
air must cease when the wind direction
becomes parallel to the isobars. Otherwise
the wind would be blowing in the direction
of increasing pressure, which would be
like water running uphill. Since the
Coriolis force acts in a direction 90
degrees to the right of the wind, it must act
in a direction opposite to the pressure
Figure 4.1.15 Illustration of the gradient wind; Ugr R, gradient at the time of maximum
radius of curvature deflection. If there are no other forces
present, this Coriolis force will exactly
balance the pressure gradient force and the wind will flow parallel to the isobars, with higher pressure
to the right of the wind direction. For straight or slightly curved isobars this resultant wind is called the
geotropic wind.
When strongly curved isobars are found, a centrifugal force must also be considered. Figure
4.1.12shows one isobar around a cyclone, which is a low pressure area rotating counter clockwise
(Northern Hemisphere). Figure 4.1.13 shows an isobar around a high pressure area which is rotating
clockwise (Northern Hemisphere). This region is called an anticyclone. As mentioned earlier, the low
pressure area is usually associated with bad weather, but does not imply anything about the magnitude
of the wind speeds. A cyclone normally covers a major part of a state or several states and has rather
gentle winds. It should not be confused with a tornado, which covers a very small region and has very
destructive winds.
The wind moving counter clockwise in the cyclone experiences a pressure gradient force fp
inward, a Coriolis force fc outward, and a centrifugal force fg outward. For wind to continue moving in
a counter clockwise direction parallel to the isobars, the forces must be balanced, so the pressure

| Wind Characteristics Pavesi - 39 -


gradient force for a cyclone is
fp = fc + fg (5.1.24)
The pressure force inward is balanced by the sum of the Coriolis and centrifugal forces. The
wind that flows in such a system is called the gradient wind. The term “geostrophic wind” applies only
to a wind flowing in the vicinity of nearly straight isobars. For the high pressure area of Figure 4.1.13,
the pressure and Coriolis forces reverse in direction. The pressure gradient force for an anticyclone is
therefore
fp = fc − fg (5.1.25)
The difference between Eqs. (5.1.24) and (5.1.25) means that cyclones and anticyclones tend to
stabilize at somewhat different relative pressures and wind speeds. Since the atmosphere is never
completely stable, these differences are not usually of major concern.
The pressure force on the air (per unit mass), fp is given by:
1 p
fp   (5.1.26)
 n
where  the density of the air and n is the direction normal to lines of constant pressure.
Also, p n is defined as the pressure gradient normal to the lines of constant pressure, or
isobars. The Coriolis force (per unit mass), fc, a fictitious force caused by measurements with respect to
a rotating reference frame (the earth), is expressed as:
fc =fU (5.1.27)
where U is the wind speed, and f is the Coriolis parameter [ f = 2w sin()].  represents the latitude
and  the angular rotation of the earth. Thus, the magnitude of the Coriolis force depends on wind
speed and latitude. The direction of the Coriolis force is perpendicular to the direction of motion of the
air. The result of these two forces, called the geostrophic wind, tends to be parallel to isobars (see
Figure 4.1.14).
The magnitude of the geostrophic wind, Ug is a function of the balance of forces and is given by:
1 p
Ug  
f n (5.1.28)
This is an idealized case since the presence of areas of high and low pressures cause the isobars to
be curved. This imposes a further force on the wind, a centrifugal force. The resulting wind, called a
gradient wind, Ugr, is shown in Figure 4.1.15.
The gradient wind is also parallel to the isobars and is the result of the balance of the forces:
Ugr2 1 p
 fUgr 
R  n (5.1.29)
where R is the radius of curvature of the path of the air particles, and
Ugr2
Ugr  Ug 
fR (5.1.30)

| Wind Characteristics Pavesi - 40 -


A final force on the wind is due to friction at the earth’s surface. That is, the earth’s surface exerts
a horizontal force upon the moving air, the effect of which is to retard the flow. This force decreases as
the height above the ground increases and becomes negligible above the boundary layer (defined as the
near earth region of the atmosphere where viscous forces are important). Above the boundary layer, a
frictionless wind balance is established, and the wind flows with the gradient wind velocity along the
isobars. Friction at the surface causes the wind to be diverted more toward the low-pressure region.
More details concerning the earth’s boundary layer and its characteristics will be given in later sections.

| Wind Characteristics Pavesi - 41 -


4.2 World Distribution of Wind
Ever since the days of sailing ships, it has been recognized that some areas of the earth’s surface
have higher wind speeds than others. Terms like doldrums, horse latitudes, and trade winds are well
established in literature. A very general picture of prevailing winds over the surface of the earth is
shown in Figure 4.2.1. In some large areas or at some seasons, the actual pattern differs strongly from
this idealized picture. These variations are due primarily to the irregular heating of the earth’s surface
in both time and position.
The equatorial calms or doldrums are due to a belt of low pressure which surrounds the earth in
the equatorial zone as a result of the average overheating of the earth in this region.
The warm air here rises in a strong convection flow. Late afternoon showers are common from
the resulting adiabatic cooling which is most pronounced at the time of highest diurnal (daily)
temperature. These showers keep the humidity very high without providing much surface cooling. The
atmosphere tends to be oppressive, hot, and sticky with calm winds and slick glassy seas. Unless
prominent land features change the weather patterns, regions near the equator will not be very good for
wind power applications.
Ideally, there are two belts of high pressure and relatively light winds which occur symmetrically
around the equator at 30° N and 30° S latitude. These are called the subtropical calms or subtropical
highs or horse latitudes. The latter name apparently dates back to the sailing vessel days when horses
were thrown overboard from becalmed ships to lighten the load and conserve water. The high pressure
pattern is maintained by vertically descending air inside the pattern. This air is warmed adiabatically
and therefore develops a low relative humidity with clear skies. The dryness of this descending air is
responsible for the bulk of the world’s great deserts which lie in the horse latitudes.
There are then two more belts of low pressure which occur at perhaps 60° S latitude and 60° N
latitude, the subpolar lows. In the Southern Hemisphere, this low is fairly stable and does not change
much from summer to winter. This would be expected because of the global encirclement by the
southern oceans at these latitudes. In the Northern Hemisphere, however, there are large land masses
and strong temperature differences between land and water. These cause the lows to reverse and
become highs over land in the winter (the Canadian and Siberian highs). At the same time the lows

Figure 4.2.1 Terrestrial pressure and wind systems

| Wind Characteristics Pavesi - 42 -


over the oceans, called the Iceland low and the Aleutian low, become especially intense and stormy low
pressure areas over the relatively warm North Atlantic and North Pacific Oceans.
Finally, the polar regions tend to be high pressure areas more than low pressure. The intensities
and locations of these highs may vary widely, with the centre of the high only rarely located at the
geographic pole.
The combination of these high and low pressure areas with the Coriolis force produces the
prevailing winds shown in Figure 4.2.1. The northeast and southeast trade winds are among the most
constant winds on earth, at least over the oceans. This causes some islands, such as Hawaii (20° N.
Latitude) and Puerto Rico (18° N. Latitude), to have excellent wind resources. The westerlies are well
defined over the Southern Hemisphere because of lack of land masses. Wind speeds are quite steady
and strong during the year, with an average speed of 8 to 14 m/s.
The wind speeds tend to increase with increasing southerly latitude, leading to the descriptive
terms roaring forties, furious fifties, and screaming sixties. This means that islands in these latitudes,
such as New Zealand, should be prime candidates for wind power sites.
In the Northern Hemisphere, the westerlies are quite variable and may be masked or completely
reversed by more prominent circulation about moving low and high pressure areas. This is particularly
true over the large land masses.

Figure 4.2.2 a, b Zonal winds (at longitude zero) for January and April of 1997, as functions of
latitude and pressure (height) (based on NCEP-NCAR, 1998).

Figure 4.2.3 a, b Zonal winds (at longitude zero) for July and October of 1997, as functions of
latitude and pressure (height) (based on NCEP-NCAR, 1998).

| Wind Characteristics Pavesi - 43 -


For two seasons, Figure 4.2.2 shows the longitude-averaged component of the large-scale wind
velocity along the x-axis, i.e. parallel to the latitude circles. This component is called the zonal wind.
At low latitudes, the zonal wind is directed towards the west, at mid-latitudes towards the east
(reaching high speeds around 12 km altitude), and again towards the west near the poles.
In Figure 4.2.3 the meridional circulation is exhibited, again for two different seasons and
averaged over longitudes. The figure gives the streamfunction, related to the mass flux ρ v* by
  , z    v*nrs cos dds
,z (5.1.31)
 2rs cos     v* nds
0

where n is normal to the area element dλ ds, and the


integration path s in the (y, z) plane connects the surface
Ψ=0 with the point (φ, z) considered. “[ ]” denotes longitude
average. The sign of Ψ along a closed loop streamline is
positive if the direction of flow is towards south at the
ground and towards north aloft. The loops are called
meridional cell motion, and one notes a cell of large mass
transport near the Equator. Its direction changes with
season.
On an average annual basis, there are two additional
cells on each side of the Equator. They are called Hadley
Figure 4.2.4 Summer (top) and winter cells (Hadley, 1735), the northern one being negative and
(bottom) streamfunctions for the the southern positive, according to the above sign
meridional mass transport (longitude convention. Near the poles there is an even weaker set of
average), given in units of 109 kg s-2. meridional cells without seasonal sign changes. 1997 data
Cells with a positive streamfunction has indicate that all the cells extend up to about 200 mb, in
northward transport at top (based on contrast to Figure 4.2.3 (NASA, 1998). The cell structure is
Newell et al., 1969). not entirely stable and features transient cells of shorter
duration.
Combining the standing cell picture with the mean
zonal winds, one observes the following regularities near the
Earth’s surface: In a region near the Equator, the westward
zonal winds combine with meridional transport towards the
Equator (from both sides) to form the trade winds (blowing
towards the south-west in the northern hemisphere and
Figure 4.2.5 Northward transport of towards north-west in the southern hemisphere). At mid-
angular momentum by eddies, averaged latitudes the eastward zonal winds combine with meridional
over year and longitudes and expressed as cell motion away from the Equator, to produce the
1018 kg m2s-1 per unit of pressure layer, "westerlies” (blowing towards the north-east in the northern
i.e. per 104 N m-2. Expressed per unit of hemisphere and towards south-east in the southern
height z, the magnitude of transport would hemisphere). Further away from the Equator, in the polar
increase less rapidly upwards towards the regions, the wind directions are as for the trade winds.
two maxima (based on Buch, 1954; Obasi, The deviation of the wind field from the longitude
1963; also using Lorenz, 1967). average is not unimportant. A part of it is quite regular, in

| Wind Characteristics Pavesi - 44 -


the form of standing horizontal cells, which are usually denoted (large-scale) eddies. They play an
important role in transporting angular momentum towards the poles. Figure 4.2.4 shows the annual and
longitude average of this transport.
Angular momentum is added to the atmosphere due to friction and drag forces between the
atmosphere and the Earth’s surface. Figure 4.2.5 gives a sketch of the torque acting on the atmosphere
as a result of these forces. The main part of the torque is due to friction, but the contribution from drag
forces is not negligible in mountain regions. The figure is based on a rough estimate by Priestley
(1951), but it is consistent with the angular momentum transport in the atmosphere, as mentioned by
Lorenz (1967).
The angular momentum added to the atmosphere at low latitudes is transported to a height of
about 10 km by the meridional cell motion. Here it is taken over by the large-scale horizontal eddy
motion, which brings it to mid-latitudes. Practically no transport across latitudes is performed by the
meridional cell motion itself, as can be seen from Figure 4.2.5. The absolute angular momentum (i.e.
measured in a coordinate system not following the Earth’s rotation) of a certain volume of the
atmosphere is proportional to the tangential velocity and to the distance to the Earth’s axis of rotation,
rcosφ. If the same volume of air is moved polewards, the value of rcosφ decreases, and the volume of
air must either increase its tangential velocity or get rid of some of its angular momentum. Both happen
in the mid-latitude atmosphere.
The regions of increased tangential velocity are easily recognised at a height of about 12 km and
latitudes of 30–50° (pressure about 200 mb, cf. Figure 4.2.5). Some angular momentum is lost by
internal friction in the atmosphere, and some is transported back down to Earth level, where it
compensates for the angular momentum losses, occurring as a result of friction with the surface. The
fact that the rotation of the Earth is not accelerating or decelerating shows that, on average, friction
forces in opposite direction cancel each other out at low and mid-latitudes. At latitudes above roughly
60°, the transport of angular momentum reverses sign, becoming a net transport towards the Equator. It
has been pointed out by Lorenz (1967) that the transport of angular momentum by large eddies cannot
be regarded as a simple diffusion process, since the corresponding “eddy diffusion coefficient” k would
in that case have to be negative over a large part of the atmosphere. Indeed the transport of angular
momentum has to be seen as an integral part of the general circulation, which cannot be considered to
consist of just zonal and meridional motion. This is also inherent in the equation of motion written in
terms of density averaged quantities, which contains the horizontal eddy motion as a combination of
the velocity vector components Vx and Vy

4.2.1 Other atmospheric circulation patterns


The general circulation flow pattern described previously best represents a model for a smooth
spherical surface. In reality, the earth's surface varies considerably, with large ocean and land masses.
These different surfaces can affect the flow of air due to variations in pressure fields, the absorption of
solar radiation, and the amount of moisture available. The oceans act as a large sink for energy.
Therefore, the movement of air is often coupled with the ocean circulation. All these effects lead to
differential pressures which affect the global winds and many of the persistent regional winds, such as
occur during monsoons. In addition, local heating or cooling may cause persistent local winds to occur
on a seasonal or daily basis. These include sea breezes and mountain winds.
Smaller scale atmospheric circulation can be divided into secondary and tertiary circulation (see
Rohatgi and Nelson, 1994). Secondary circulation occurs if the centres of high or low pressure are
caused by heating or cooling of the lower atmosphere. Secondary circulations include the following:

| Wind Characteristics Pavesi - 45 -


 Hurricanes
 Monsoon circulation
 Extratropical cyclones
Tertiary circulations are small-scale, local circulations characterized by local winds. These
include the following:
 Land and sea breezes
 Valley and mountain winds
 Monsoon-like flow (example: flow in California passes)
 Foehn winds (dry high-temperature winds on the downwind side of mountain ranges)
 Thunderstorms
 Tornadoes
Changes in velocity and direction of wind near the surface, say up to 100 m above the ground, is
more important as far as energy conversion is concerned. In this region, the wind pattern is further
influenced by several local factors.
Land and sea breezes are examples for the local wind effects. During the day time, land gets
heated faster than the sea surface. As a result, the air near the land rises, forming a low pressure region.
This attracts cool air to the land from the sea. This is called the sea breeze. During night time, the
process gets reversed as
cooling is faster on land.
Thus wind blows from the
land to the sea, which is
called the land breeze.
In mountain valleys, the
air above the surface gets
heated and rises up along the
slopes during the day time.
This is replaced by the cool Figure 4.2.6 Diurnal valley and mountain wind
air, resulting in the valley
winds. During the night, the
flow is from the mountain to
the valley which is known as
the mountain wind. Quite
often, this phenomenon may
create very strong air
currents, developing
powerful wind. Wind shear,
turbulence and acceleration
over the ridges are some
other examples for local wind
effects.
Examples of tertiary
circulation, valley and
mountain winds, are shown
in Figure 4.2.6. During the Figure 4.2.7 Time and space scales of atmospheric motion (Spera,
day, the warmer air of the 1994).

| Wind Characteristics Pavesi - 46 -


mountain slope rises and replaces the heavier cool air above it. The direction reverses at night, as cold
air drains down the slopes and stagnates in the valley floor. An understanding of these wind patterns,
and other local effects is important for the evaluation of potential wind energy sites.
Atmospheric motions vary in both time (seconds to months) and space (centimetres to thousands
of kilometres). Figure 4.2.7 summarizes the time and space variations of atmospheric motion as applied
to wind energy. As will be discussed in later sections, space variations are generally dependent on
height above the ground and global and local geographical conditions.

4.2.2 Variations in time


Following conventional practice, variations of wind speed in time can be divided into the
following categories:
 Inter-annual
 Annual
 Diurnal
 Short-term (gusts and turbulence)
A review of each of these categories as well as comments on wind speed variation due to location
and wind direction follows.

4.2.2.1 Inter-annual

Inter-annual variations in wind speed occur over time scales greater than one year. They can have
a large effect on long-term wind turbine production. The ability to estimate the inter-annual variability
at a given site is almost as important as estimating the long-term mean wind at a site. Meteorologists
generally conclude that it takes 30 years
of data to determine long-term values of
weather or climate and that it takes at
least five years to arrive at a reliable
average annual wind speed at a given
location. Aspliden et al. (1986) note
that one statistically developed rule of
thumb is that one year of record data is
generally sufficient to predict long-term
seasonal mean wind speeds within an
accuracy of 10% with a confidence
level of 90%.
Researchers are still looking for
reliable prediction models for long-term
mean wind speed. The complexities of
the interactions of the meteorological Figure 4.2.8 Seasonal changes of monthly average wind
and topographical factors that cause its speeds (Hiester and Pennell, 1981)
variation make the task difficult.

4.2.2.2 Annual

Significant variations in seasonal or monthly averaged wind speeds are common over most of the

| Wind Characteristics Pavesi - 47 -


world.
Figure 4.2.8 illustrates seasonal changes of monthly wind speed. It is interesting to note that this
figure clearly shows that the typical behaviour of monthly variation is not defined by a single year of
data.
Similarly, Figure 4.2.9 provides an illustration of the importance of annual wind speed variation,
and its effect on available wind power.

4.2.2.3 Diurnal (time of day)

In both tropical and temperate


latitudes, large wind variations also can
occur on a diurnal or daily time scale. This
type of wind speed variation is due to
differential heating of the earth’s surface
during the daily radiation cycle. A typical
diurnal variation is an increase in wind
speed during the day with the wind speeds
lowest during the hours from midnight to
sunrise. Daily variations in solar radiation
are responsible for diurnal wind variations
in temperate latitudes over relatively flat Figure 4.2.9 Seasonal variation of available wind
land areas. The largest diurnal changes power per unit area (Rohatgi and Nelson, 1994).
generally occur in spring and summer, and
the smallest in winter.
Furthermore, the diurnal variation in
wind speed may vary with location and
altitude above sea level. For example, at
altitudes high above surrounding terrain,
e.g., mountains or ridges, the diurnal pattern
may be very different. This variation can be
explained by mixing or transfer of
momentum from the upper air to the lower
air.
As illustrated in Figure 4.2.10, there
may be significant year-to-year differences
in diurnal behaviour, even at fairly windy
locations. Although gross features of the
diurnal cycle can be established with a
single year of data, more detailed features
such as the amplitude of the diurnal
oscillation and the time of day that the
maximum winds occur cannot be
determined precisely.
Figure 4.2.10 Monthly mean diurnal wind speeds for
January and July (Hiester and Pennell, 1981)

| Wind Characteristics Pavesi - 48 -


4.2.2.4 Short-term

Short-term wind speed variations of interest include turbulence and gusts.


Figure 4.2.11, output from an anemometer (described later), shows the type of short-term wind
speed variations that normally exist. Short-term variations usually mean variations over time intervals
of 10 minutes or less. Ten-minute averages are typically determined using a sampling rate of about 1
second. It is generally accepted that variations
in wind speed with periods from less than a
second to 10 minutes and that have a
stochastic character are considered to represent
turbulence. For wind energy applications,
turbulent fluctuations in the flow need to be
quantified for the turbine design considerations
based on maximum load and fatigue
prediction, structural excitations, control.
system operation, and power quality. More
details on these factors as related to turbine
Figure 4.2.11 Typical plot of wind speed vs. time for a design are discussed in next chapters.
short period Turbulence can be thought of as random
wind speed fluctuations imposed on the mean
wind speed. These fluctuations occur in all
three directions: longitudinal (in the direction
of the wind), lateral (perpendicular to the
average wind), and vertical. A gust is a
discrete event within a turbulent wind field. As
illustrated in Figure 4.2.12, one way to
characterize the attributes of a gust is to
measure: (a) amplitude, (b) rise time, (c)
maximum gust variation, and (d) lapse time.
Wind turbine loads caused by gusts can be
determined as functions of these four
Figure 4.2.12 Illustration of a discrete gust event; a,
attributes. For example, extreme loads can be
amplitude; b, rise time; c, maximum gust variation; d,
analyzed by determining the response to the
lapse time
largest gust during a wind machine’s expected
lifetime.

4.2.2.5 Variation due to the location

Wind speed is also very dependent on


local topographical ground cover variations.
For example as shown in Figure 4.2.13
(Hiester and Pennell. 1981), differences
between two sites close to each other can be
significant. The graph shows monthly and
Figure 4.2.13 Time series of monthly wind speeds five-year mean wind speeds for two sites 21
(Hiester and Pennell, 1981)

| Wind Characteristics Pavesi - 49 -


km apart.
The five-year average mean wind speeds differ by about 12% (4.75 and 4.25 m/s annual
averages).

4.2.3 Variation due to wind direction


Wind direction also varies over the same time scales over which wind speeds vary. Seasonal
variations may be small, on the order of 30 degrees, or the average monthly winds may change
direction by 180 degrees over a year. Short-term direction variations are the result of the turbulent
nature of the wind. These short-term variations in wind direction need to be considered in wind turbine
design and sitting.
Horizontal axis wind turbines must rotate (yaw) with changes in wind direction. Yawing causes
gyroscopic loads throughout the turbine structure and exercises any mechanism involved in the yawing
motion. Crosswind due to changes in wind direction affect blade loads. Thus, short term variations in
wind direction and the associated motion affect the fatigue life of components such as blades and yaw
drives.

| Wind Characteristics Pavesi - 50 -


4.3 Equation Section (Next)Atmospheric Stability
As we have mentioned, most wind speed measurements are made about 10 m above the ground.
Small wind turbines are typically mounted 20 to 30 m above ground level, while the propeller tip may
reach a height of more than 100 m on the large turbines, so an estimate of wind speed variation with
height is needed. The mathematical form of this variation will be considered in the next section, but
first we need to examine a property of the atmosphere which controls this vertical variation of wind
speed. This property is called atmospheric stability, which refers to the amount of mixing present in the
atmosphere. We start this discussion by examining the pressure variation with height in the lower
atmosphere
A given parcel of air has mass and is attracted to the earth by the force of gravity. For the parcel
not to fall to the earth’s surface, there must be an equal and opposite force directed away from the
earth. This force comes from the decrease in air pressure with increasing height. The greater the density
of air, the more rapidly must pressure decrease upward to hold the air at constant height against the
force of gravity. Therefore, pressure decreases quickly with height at low altitudes, where density is
high, and slowly at high altitudes where density is low. This balanced force condition is called
hydrostatic balance or hydrostatic equilibrium.
The average atmospheric pressure as a function of elevation above sea level for middle latitudes
is shown in Figure 4.3.1. This curve is part of the model for the U.S. Standard Atmosphere. At sea level
and a temperature of 273 K, the average pressure is 101.3 kPa, as mentioned earlier. A pressure of half
this value is reached at about 5500 m. This pressure change with elevation is especially noticeable
when flying in an airplane, when one’s ears tend to ‘pop’ as the airplane changes altitude.
It should be noticed that the independent variable z is plotted on the vertical axis in this figure,
while the dependent variable is plotted along the horizontal axis. It is plotted this way because
elevation is intuitively up. To read the graph
when the average pressure at a given height is
desired, just enter the graph at the specified
height, proceed horizontally until you hit the
curve, and then go vertically downward to
read the value of pressure.
The pressure in Figure 4.3.1 is assumed
to not vary with local temperature. That is, it
is assumed that the column of air directly
above the point of interest maintains the same
mass throughout any temperature cycle. A
given mass of air above some point will
produce a constant pressure regardless of the
temperature. The volume of gas in the column
will change with temperature but not the
pressure exerted by the column. This is not a
perfect assumption because, while the mass of
the entire atmosphere does not vary with
temperature, the mass directly overhead will
Figure 4.3.1 Pressure variation with altitude for U.S. vary somewhat with temperature. A
Standard Atmosphere temperature decrease of 30°C will often be

| Wind Characteristics Pavesi - 51 -


associated with a pressure increase of 2 to 3 kPa. The atmospheric pressure tends to be a little higher in
the early morning than in the middle of the afternoon. Winter pressures tend to be higher than summer
pressures. This effect is smaller than the pressure variation due to movement of weather patterns, hence
will be ignored in this text.
It will be seen later that the power output of a wind turbine is proportional to air density, which in
turn is proportional to air pressure. A given wind speed therefore produces less power from a particular
turbine at higher elevations, because the air pressure is less. A wind turbine located at an elevation of
1000 m above sea level will produce only about 90 % of the power it would produce at sea level, for
the same wind speed and air temperature. Therefore this pressure variation with elevation must be
considered in both technical and economic studies of wind power.
The air density at a proposed wind turbine site is estimated by finding the average pressure at that
elevation from Figure 4.3.1 and then using Eq. (5.1.6) to find density. The ambient temperature must be
used in this equation. We have seen that the average pressure at a given site is a function of elevation
above sea level. Ground level temperatures vary in a minor way with elevation, but are dominated by
latitude and topographic features. We must be cautious about estimating ground level temperatures
based on elevation above sea level. Once we leave the ground, however, and enter the first few
thousand meters of the atmosphere, we find a more predictable temperature decrease with altitude. We
shall use the word altitude to refer to the height of an object above ground level, and elevation to refer
to the height above sea level. With these definitions, a pilot flying in the mountains is much more
concerned about his altitude than his elevation.
This temperature variation with altitude is very important to the character of the winds in the first
200 m above the earth’s surface, so we shall examine it in some detail. We observe that Eq. (5.1.3) (the
ideal gas law or the equation of state) can be satisfied by pressure and temperature decreasing with
altitude while the volume of one kmol (the specific volume) is increasing.
Pressure and temperature are easily measured while specific volume is not. It is therefore
common to use the first law of thermodynamics (which states that energy is conserved) and eliminate
the volume term from Eq. (5.1.3). When this is done for an adiabatic process , the result is:
R
T1  p1  cp
  (5.3.1)
T0  p0 
where T1 and p1 are the temperature and pressure at state 1, T0 and p0 are the temperature and
pressure at state 0, R is the universal gas constant, and cp is the constant-pressure specific heat of air.
The derivation of this equation may be found in most introductory thermodynamics books. The average
value for the ratio R/cp in the lower atmosphere is 0.286.
Eq. (5.3.1), sometimes called Poisson’s equation, relates adiabatic temperature changes
experienced by a parcel of air undergoing vertical displacement to the pressure field through which it
moves. If we know the initial conditions T0 and p0, we can calculate the temperature T1 at any pressure
p1 as long as the process is adiabatic and involves only ideal gases.
Ideal gases can contain no liquid or solid material and still be ideal. Air behaves like an ideal gas
as long as the water vapour in it is not saturated. When saturation occurs, water starts to condense, and
in condensing gives up its latent heat of vaporization. This heat energy input violates the adiabatic
constraint, while the presence of liquid water keeps air from being an ideal gas.
We see from this example that a parcel of air which undergoes an upward displacement
experiences a temperature decrease of about 1oC/100 m in an adiabatic process. This quantity is

| Wind Characteristics Pavesi - 52 -


referred to as the adiabatic lapse rate, the dry- adiabatic lapse rate, or the temperature gradient. It is an
ideal quantity, and does not vary with actual atmospheric parameters. This ideal temperature decrease
is reasonably linear up to several kilometres above the earth’s surface and can be approximated by
Ta(z) = Tg − Ra(z − zg) (5.3.2)
where Ta(z) is the temperature at elevation z m above sea level if all processes are adiabatic, Tg is
the temperature at ground level zg, and Ra is the adiabatic lapse rate, 0.01°C/m. The quantity z − zg is
the altitude above ground level.
The actual temperature decrease with height will normally be different from the adiabatic
prediction, due to mechanical mixing of the atmosphere. The actual temperature may decrease more
rapidly or less rapidly than the adiabatic decrease. In fact, the actual temperature can even increase for
some vertical intervals. A plot of some commonly observed temperature variations with height is
shown in Figure 4.3.2. We shall start the explanation at 3 p.m., at which time the earth has normally
reached its maximum temperature and the first 1000 m or so above the earth’s surface is well mixed.
This means that the curve of actual temperature will follow the theoretical adiabatic curve rather
closely. By 6 p.m. the ground temperature has normally dropped slightly. The earth is much more
effective at receiving energy from the sun and
reradiating it into space than is the air above the
earth. This means that the air above the earth is
cooled and heated by conduction and convection
from the earth, hence the air temperature tends to
lag behind the ground temperature. This is shown
in Figure 4.3.2b, where at 6 p.m. the air
temperature is slightly above the adiabatic line
starting from the current ground temperature. As
ground temperature continues to drop during the
night, the difference between the actual or
prevailing temperature and the adiabatic curve
becomes even more pronounced. There may even
be a temperature inversion near the ground. This
usually occurs only with clear skies, which allow
the earth to radiate its energy into space effectively.
When sunlight strikes the earth in the
morning, the earth’s surface temperature rises
rapidly, producing the air temperature variations
shown in Figure 4.3.2a. The difference between the
actual air temperature and the adiabatic curve
would normally reach its maximum before noon,
causing a relatively rapid heating of the air. There
will be a strong convective flow of air during this
time because a parcel of air that is displaced
upward will find itself warmer and hence lighter
than its surroundings. It is then accelerated
Figure 4.3.2 Dry-adiabatic and actual upwards under hydrostatic pressure.
temperature variations with height as a function of It will continue to rise until its temperature is
time of day, for clear skies.

| Wind Characteristics Pavesi - 53 -


the same as that of the adiabatic curve. Another parcel has to move down to take the first parcel’s
place, perhaps coming down where the earth is not as effective as a collector of solar energy, and hence
causes the atmosphere to be well mixed under these conditions. This condition is referred to as an
unstable atmosphere.
On clear nights, however, the earth will be colder than the air above it, so a parcel at the
temperature of the earth that is displaced upward will find itself colder than the surrounding air. This
makes it more dense than its surroundings so that it tends to sink back down to its original position.
This condition is referred to as a stable atmosphere. Atmospheres which have temperature profiles
between those for unstable and stable atmospheres are referred to as neutral atmospheres. The daily
variation in atmospheric stability and surface wind speeds is called the diurnal cycle.
It is occasionally convenient to express the actual temperature variation in an equation similar to
Eq. (5.3.2). Over at least a narrow range of heights, the prevailing temperature Tp(z) can be written as
Tp(z) = Tg − Rp(z − zg) (5.3.3)
where Rp is the slope of a straight line approximation to the actual temperature curve called the
prevailing lapse rate. Tg is the temperature at ground level, zg m above mean sea level, and z is the
elevation of the observation point above sea level. We can
force this equation to fit one of the dashed curves of Figure
4.3.2 by using a least squares technique, and determine an
approximate lapse rate for that particular time. If we do this
for all times of the day and all seasons of the year, we find
that the average prevailing lapse rate Rp is 0.0065° C/m.
Suppose that a parcel of air is heated above the
temperature of the neighbouring air so it is now buoyant and
will start to rise. If the prevailing lapse rate is less than
adiabatic, the parcel will rise until its temperature is the same
as the surrounding air. The pressure force and hence the
acceleration of the parcel is zero at the point where the two
lapse rate lines intersect. The upward velocity, however,
Figure 4.3.3 Buoyancy of air in a stable produced by acceleration from the ground to the height at
atmosphere which the buoyancy vanishes, is greatest at that point. Hence
the air will continue upward, now colder and more dense
than its surroundings, and decelerate. Soon the upward
motion will cease and the parcel will start to sink. After a few
oscillations about that level the parcel will settle near that
height as it is slowed down by friction with the surrounding
air.
The two temperatures are the same at the point of equal
buoyancy. Setting Ta equal to Tp and solving for z yields an
intersection height of about 585 m above sea level or 285 m
above ground level. This is illustrated in Figure 4.3.3.
Atmospheric mixing may be limited to a relatively shallow
layer if there is a temperature inversion at the top of the
layer. This situation is illustrated in Figure 4.3.4. Heated
Figure 4.3.4 Deep buoyant layer topped
parcels rise until their temperature is the same as that of the
by temperature inversion.

| Wind Characteristics Pavesi - 54 -


ambient air, as before. Now, however, a doubling of the initial temperature difference does not result in
a doubling of the height of the intersection, but rather yields a minor increase. Temperature inversions
act as very strong lids against penetration of air from below, trapping the surface air layer underneath
the inversion base.
A more detailed system of defining stability is especially important in atmospheric pollution
studies. Various stability parameters or categories have been defined and are available in the literature
[Pasquill, F. (1974)].
A stable atmosphere may have abrupt changes in wind speed at a boundary layer. The winds may
be nearly calm up to an elevation of 50 or 100 m, and may be 20 m/s above that boundary. A horizontal
axis wind turbine which happened to have its hub at this boundary would experience very strong
bending moments on its blades and may have to be shut down in such an environment. An unstable
atmosphere will be better mixed and will not evidence such sharp boundaries.
The wind associated with an unstable atmosphere will tend to be gusty, because of thermal
mixing. Wind speeds will be quite small near the ground because of ground friction, increasing upward
for perhaps several hundred meters over flat terrain. A parcel of heated air therefore leaves the ground
with a low horizontal wind speed. As it travels upward, the surrounding air exerts drag forces on the
parcel, tending to speed it up. The parcel will still maintain its identity for some time, travelling slower
than the surrounding air.
Parcels of air descending from above have higher horizontal speeds. They mix with the ascending
parcels, causing an observer near the ground to sense wind speeds both below and above the mean
wind speed. A parcel with higher velocity is called a gust, while a parcel with lower velocity is
sometimes called a negative gust or a lull. These parcels vary widely in size and will hit a particular
point in a random fashion. They are easily observed in the wave structure of a lake or in a field of tall
wheat.
Gusts pose a hazard to large wind turbines because of the sudden change in wind speed and
direction experienced by the turbine during a gust. Vibrations may be established or structural damage
done. The wind turbine must be designed to withstand the peak gust that it is likely to experience
during its projected lifetime.
Gusts also pose a problem in the adjusting of the turbine blades during operation in that the
sensing anemometer will experience a different wind speed from that experienced by the blades. A gust
may hit the anemometer and cause the blade angle of attack (the angle at which the blade passes
through the air) to be adjusted for a higher wind speed than the blades are actually experiencing. This
will generally cause the power output to drop below the optimum amount. Conversely, when the
sensing anemometer experiences a lull, the turbine may be trimmed (adjusted) to produce more than
rated output because of the higher wind speed it is experiencing. These undesirable situations require
the turbine control system to be rather complex, and to have long time constants. This means that the
turbine will be aimed slightly off the instantaneous wind direction and not be optimally adjusted for the
instantaneous wind speed a large fraction of the time. The power production of the turbine will
therefore be somewhat lower than would be predicted for an optimally adjusted and aimed wind turbine
in a steady wind. The amount of this reduction is difficult to measure or even to estimate, but could
easily be on the order of a 10% reduction. This makes the simpler vertical axis machines more
competitive than they might appear from wind tunnel tests since they require no aiming or blade
control.
Thus far, we have talked only about gusts due to thermal turbulence, which has a very strong
diurnal cycle. Gusts are also produced by mechanical turbulence, caused by higher speed winds

| Wind Characteristics Pavesi - 55 -


flowing over rough surfaces. When strong
frontal systems pass through a region, the
atmosphere will be mechanically mixed and
little or no diurnal variation of wind speed
will be observed. When there is no
significant mixing of the atmosphere due to
either thermal or mechanical turbulence, a
boundary layer may develop with relatively
high speed laminar flow of air above the
boundary and essentially calm conditions
below it. Without the effects of mixing,
upper level wind speeds tend to be higher
than when mixed with lower level winds.
Thus it is quite possible for nighttimes
wind speeds to decrease near the ground and
increase a few hundred meters above the
ground. This phenomenon is called the
nocturnal jet. As mentioned earlier, most
National Weather Service (NWS)
anemometers have been located about 10 m
above the ground, so the height and
frequency of occurrence of this nocturnal jet
have not been well documented at many
sites. Investigation of nocturnal jets is done
either with very tall towers (e.g. 200 m) or
with meteorological balloons. Balloon data
Figure 4.3.5 Western Kansas wind data, 1970-1973.
are not very precise, as we shall see in more
detail in the next chapter, but rather long term records are available of National Weather Service
balloon launchings. When used with appropriate caution, these data can show some very interesting
variations of wind speed with height and time of day.
Figure 4.3.5 shows the diurnal wind speed pattern at Dodge City, Kansas for the four year period,
1970-73, as recorded by the National Weather Service with an anemometer at 7 m above the ground.
The average wind speed for this period was 5.66 m/s at this height. The spring season (March, April,
and May) is seen to have the highest winds, with the summer slightly lower than the fall and winter
seasons. The wind speed at 7 m is seen to have its peak during the middle of the day for all seasons.
Also shown in Figure 4.3.5 are the results from two balloon launchings per day for the same period for
three Kansas locations, Dodge City, Goodland, and Wichita. The terrain and wind characteristics are
quite similar at each location, and averaging reduces the concern about missing data or local anomalies.
The surface wind speed is measured at the time of the launch. These values lie very close to the Dodge
City curves, indicating reasonable consistency of data. The wind speed pattern indicated by the
balloons at 216 m is then seen to be much different from the surface speeds. The diurnal cycle at 216 m
is opposite that at the surface. The lowest readings occur at noon and the highest readings occur at
midnight. The lowest readings at 216 m are higher than the highest surface winds by perhaps 10-20 %,
while the midnight wind speeds at 216 m are double those on the surface. This means that a wind
turbine located in the nocturnal jet will have good winds in the middle of the day and even better winds

| Wind Characteristics Pavesi - 56 -


at night. The average wind speed at 216 m for this period was 9.22 m/s, a very respectable value for
wind turbine operation.
Measurements at intermediate heights will indicate some height where there is no diurnal cycle. A
site at which the wind speed averages 7 or 8 m/s is a good site, especially if it is relatively steady. This
again indicates the importance of detailed wind measurements at any proposed site. Two very similar
sites may have the same surface winds, but one may have a well developed nocturnal jet at 20 m while
the other may never have a nocturnal jet below 200 m. The energy production of a wind turbine on a
tall tower at the first site may be nearly double that at the second site, with a corresponding decrease in
the cost of electricity. We see that measurements really need to be taken at heights up to the hub height
plus blade radius over a period of at least a year to clearly indicate the actual available wind.

| Wind Characteristics Pavesi - 57 -


4.4 Wind Speed Variation with Height
The flow of air above the ground is
retarded by frictional resistance offered by the
earth surface (boundary layer effect). This
resistance may be caused by the roughness of
the ground itself or due to vegetations,
buildings and other structures present over the
ground. For example, a typical vertical wind
profile at a site is shown in Figure 4.4.1.
Theoretically, the velocity of wind right over
the ground surface should be zero. Velocity
increases with height up to a certain elevation.
In the above example, the velocity increases
noticeably up to 20 m, above which the
surface influence is rather feeble.
The rate at which the velocity increases
with height depends on the roughness of the
terrain. Presence of dense vegetations like
Figure 4.4.1 Variation of wind velocity with height plantations, forests, and bushes slows down
the wind considerably. Level and smooth
terrains do not have much effect on the wind
speed. The surface roughness of a terrain is
usually represented by the roughness class or
roughness height. The roughness height of a
surface may be close to zero (surface of the
sea) or even as high as 2 (town centres). Some
typical values are 0.005 for flat and smooth
terrains, 0.025-0.1 for open grass lands, 0.2 to
0.3 for row crops, 0.5 to 1 for orchards and
shrubs and 1to 2 for forests, town centres etc.
Roughness height is an important factor
to be considered in the design of wind energy
plants. Suppose we have a wind turbine of 30
m diameter and 40 m tower height, installed
Figure 4.4.2 Velocity ratio with respect to 10 m for over the terrain described in Figure 4.4.1. The
different roughness heights tip of the blade, in its lower position, would be
25 m above the ground. Similarly, at the extreme upper position, the blade tip is 55 m above the
ground. As we see, the wind velocities at these heights are different. Thus, the forces acting on the
blades as well as the power available would significantly vary during the rotation of the blades. This
effect can be minimized by increasing the tower height. The wind data available at meteorological
stations might have been collected from different sensor heights. In most of the cases, the data are
logged at 10 m as per recommendations of the World Meteorological Organization (WMO). In wind
energy calculations, we are concerned with the velocity available at the rotor height. The data collected
at any heights can be extrapolated to other heights on the basis of the roughness height of the terrain.

| Wind Characteristics Pavesi - 58 -


Thus, if the velocity of wind measured at a height of 10 m is 7 m/s and the roughness height is
0.1, the velocity at 40 m above the ground is 9.1 m/s. It should be noted that the power available at 40
m is 2.2 times higher than at 10 m. The wind velocities at different heights relative to that at 10 m, as
affected by the roughness heights, are shown in Figure 4.4.2. In some cases, we may have data from a
reference location (meteorological station for example) at a certain height. This data is to be
transformed to a different height at another location with similar wind profile but different roughness
height (for example, the wind turbine site). Under such situations, it is logical to assume that the wind
velocity is not significantly affected by the surface characteristics beyond a certain height. This height
may be taken as 60 m from the ground level.
As we have seen, a knowledge of wind speeds at heights of 20 to 120 m above ground is very
desirable in any decision about location and type of wind turbine to be installed. Many times, these data
are not available and some estimate must be made from wind speeds measured at about 10 m. This
requires an equation which predicts the wind speed at one height in terms of the measured speed at
another, lower, height. Such equations are developed in texts on fluid mechanics. The derivations are
beyond the scope of this text so we shall just use the results. One possible form for the variation of
wind speed u(z) with height z is
uf  z  z 
u z  ln     (5.4.1)
K  z0  L 
Here uf is the friction velocity, K is the von Karmans constant (normally assumed to be 0.4), z0 is
the surface roughness length, and L is a scale factor called the Monin Obukov length [Panofsky, H. A.
(1977), Justus, C. G. (1978)]. The function ξ(z/L) is determined by the net solar radiation at the site.
This equation applies to short term (e.g. 1 minute) average wind speeds, but not to monthly or yearly
averages. The surface roughness length z0 will depend on both the size and the spacing of roughness
elements such as grass, crops, buildings, etc. Typical values of z0 are about 0.01 cm for water or snow
surfaces, 1 cm for short grass, 25 cm for tall grass or crops, and 1 to 4 m for forest and city [Putnam,
P.C. (1948)]. In practice, z0 cannot be determined precisely from the appearance of a site but is
determined from measurements of the wind. The same is true for the friction velocity uf , which is a
function of surface friction and the air density, and ξ(z/L). The parameters are found by measuring the
wind at three heights, writing Eq. (5.4.1) as three equations (one for each height), and solving for the
three unknowns uf , z0, and ξ(z/L). This is not a linear equation so nonlinear analysis must be used. The
results must be classified by the direction of the wind and the time of year because z0 varies with the
upwind surface roughness and the condition of the vegetation. Results must also be classified by the
amount of net radiation so the appropriate functional form of ξ(z/L) can be used.
All of this is quite satisfying for detailed studies on certain critical sites, but is too difficult to use
for general engineering studies. This has led many people to look for simpler expressions which will
yield satisfactory results even if they are not theoretically exact. The most common of these simpler
expressions is the power law, expressed as

u  z2   z2 

  (5.4.2)
u  z1   z1 
In this equation z1 is usually taken as the height of measurement, approximately 10 m, and z2 is
the height at which a wind speed estimate is desired. The parameter α is determined empirically. The
equation can be made to fit observed wind data reasonably well over the range of 10 to perhaps 100 or

| Wind Characteristics Pavesi - 59 -


150 m if there are no sharp boundaries in the flow. The exponent α varies with height, time of day,
season of the year, nature of the terrain, wind speeds, and temperature [Golding, E. (1976)]. A number
of models have been proposed for the variation of α with these variables [Spera, D. A., and T. R.
Richards (1979)]. We shall use the linear logarithmic plot shown in Figure 4.4.3.
This figure shows one plot for day and another plot for night, each varying with wind speed
according to the equation
α = a − b log10 u(z1) (5.4.3)
The coefficients a and b can be determined by a linear regression program. Typical values of a
and b are 0.11 and 0.061 in the daytime and 0.38 and 0.209 at night.
Several sets of this figure can be generated at each site if necessary. One figure can be developed
for each season of the year, for example. Temperature, wind direction, and height effects can also be
accommodated by separate figures. Such figures would be valid at only the site where they were
measured. They could be used at other sites with similar terrain with some caution, of course.
The average value of α has been determined by many measurements around the world to be about
one-seventh. This has led to the common reference to Eq. (5.4.2) as the one-seventh power law
equation. This average value should be used only if site specific data is not available, because of the
wide range of values that α can assume.

4.4.1 Turbulence
The speed and direction of wind change rapidly while it passes through rough surfaces and
obstructions like buildings, trees and rocks. This is due to the turbulence generated in the flow. Extent
of this turbulence at the upstream and downstream of the flow is shown in Figure 4.4.4. The presence
of turbulence in the flow not only reduces the power available in the stream, but also imposes fatigue
loads on the turbine.
Intensity of the turbulence depends
on the size and shape of the obstruction.
Based on its nature, the turbulent zone
can extend up to 2 times the height of
the obstacle in the upwind side and 10 to
20 times in the downwind side. Its
influence in the vertical direction may
be prominent to 2 to 3 times the obstacle
height.
Hence before citing the turbine, the
obstacles present in the nearby area
should be taken in to account. The tower
should be tall enough to overcome the
influence of the turbulence zone.

4.4.2 Acceleration effect


A smooth ridge, as shown in Fig.
6.4.5, accelerates the wind stream
Figure 4.4.3 Variation of wind profile exponent α with passing over it. The acceleration is
reference wind speed u(z1). caused by the squeezing of wind layers

| Wind Characteristics Pavesi - 60 -


over the mount as shown in the figure. The
degree of acceleration depends on the shape
of the ridge. This effect can be fully
exploited for energy generation, if the slope
of the ridge is between 6° and 16°. Angles
greater than 27° and less than 3° are not
favourable. Another important factor is the
orientation of the ridge. The acceleration
effect is high when the prevailing wind is
perpendicular and low when it is parallel to
the ridge line. Similarly, if the ridge has a
Figure 4.4.4 Turbulence created by an obstruction concave side facing the wind, the effect is
more desirable. Triangular shaped ridges
offer better acceleration followed by the smooth and round geometry. Flat topped ridges may pose the
problem of turbulence, especially in the lower region.
Mountain passes are another geographical feature causing acceleration of wind.
While the flow passes through the notches in the mountain barriers, due the ventury effect, the
wind velocity is enhanced. Geometrical configuration (Width, length, slope etc.) of the pass itself is the
major factor
determining the degree
of this acceleration. A
pass between two high
hills, oriented parallel
to the wind direction,
would be a cleverly
chosen site for wind
turbines. The smoother
the surface, the higher
will be the
Figure 4.4.5 The acceleration effect over ridges acceleration.

| Wind Characteristics Pavesi - 61 -


4.5 Equation Section (Next)Measurement of wind
A precise knowledge of the wind characteristics at the prospective sites is essential for the
successful planning and implementation of wind energy projects. The basic information required for
such an analysis is the speed and direction of the prevailing wind at different time scales. Ecological
factors may often be helpful in identifying a candidate site for wind power project. Wind data from the
nearby meteorological stations can give us a better understanding on the wind spectra available at the
site. However, for a precise analysis, the wind velocity and direction at the specific site has to be
measured with the help of accurate and reliable instruments.

4.5.1 Ecological indicators


Eolian features may be used as indicators for strength of the wind prevailing in an area. Eolian
features are the formations on land surface due to continuous strong wind. Sand dunes are one example
for the eolian formation. The sand particles, picked up and carried by the wind flow, are deposited back
when the wind speed is lower. The size of the particle thus carried and deposited along with the
distance can give us an indication on the average strength of the wind in that region.
0 No deformation due to wind
I Brushing
II Light flagging
III Moderate flagging
IV Strong flagging
V Flagging and clipping
VI Throwing and flagging
VII Extreme flagging
Other types of eolian features are playa lake, sediment plumes and wind scours. More information
on the use of eolian features for locating windy areas is available in [Marrs RW Marwitz J (1977)].
Another way to identify a windy site is to
observe the biological indicators. Trees and bushes
get deformed due to strong winds. The intensity and
nature of this deformation depends on the strength of
wind. This method is specifically suitable to judge the
wind in valleys, coasts and mountain terrains.
Deformations of trees due to wind effect are classified
by Putnam (1948). There are five types of
deformations under Putnam’s classification. They are
brushing, flagging, wind throwing, clipping and
carpeting. Wind brushing refers to the leeward
bending of branches and twigs of the trees.
Brushing may clearly be observed when the
trees are off their leaves. This is an indication for light
wind, which is not useful for wind energy conversion.
In flagging, the branches are stretched out to leeward,
with possible stripping of upwind branches. The range
Figure 4.5.1 Biological rating scales for the of wind speeds corresponding to the flagging effect is
wind speed of interest for energy conversion. In wind throwing,

| Wind Characteristics Pavesi - 62 -


the main trunk and branches of the tree lean away from the coming wind. This indicates the presence of
stronger wind. Under clipping, the lead branches of the tree are suppressed from growing to its normal
height due to strong wind. With extreme winds, the trees are clipped even at a very low height. This is
termed as wind carpeting.
Based on these deformations, the intensity of wind is rated on a seven point scale put forth by
Hewson and Wade (1978). These are illustrated in Figure 4.5.1, showing the top and front views of the
tree trunk. However, it should be noted that the degree of this deformation may vary from one tree
species to the other. For this reason, this method is to be calibrated with long term wind data for a given
tree variety. Once such calibration is available, the wind speed range can be directly estimated on the
basis of these biological indicators.

4.5.2 Anemometers
The indicators discussed above, along with available wind data from meteorological stations, can
give us an idea on the suitability of a given site for wind energy extraction. However, the final selection
of the site should be made on the basis of short term field measurements. Anemometers fitted on tall
masts are used for such wind measurements. Height of the mast may be the hub height of the turbine to
avoid further correction in wind speed due surface shear. As the power is sensitive to the wind
speed, good quality anemometers which are sensitive, reliable and properly calibrated should be used
for wind measurements. There are different types of anemometers. Based on the working principle,
they can be classified as:
Rotational anemometers (cup anemometers and propeller anemometers)
Pressure type anemometers (pressure tube anemometers, pressure plate anemometers and sphere
anemometers)
Thermoelectric anemometers (hot wire anemometers and hot plate anemometers)
Phase shift anemometers (ultra sonic anemometers and laser doppler anemometers)

4.5.2.1 Cup anemometer

The anemometer, most commonly used in wind energy


measurements is the cup anemometer. It consists of three (or four)
equally spaced cups attached to a centrally rotating vertical axis
through spokes (Figure 4.5.2). The cups are hemispherical or
conical in shape and made with light weight material. This is
basically a drag device. When kept in the flow, the wind exerts drag
force on the cups. The drag force is given by
1
FD  CD AV 2
2 (5.5.1)
where CD is the drag coefficient, A is the area of cup exposed
to the wind,  is the air density and V is the wind velocity.
As the drag coefficient of concave surface is more than on the
convex surface, the cup with its concave side facing the wind
experiences more drag force. This causes the cups to rotate on its
central axis. The intensity of rotation is directly proportional to the
Figure 4.5.2 Cup anemometer velocity of incoming wind.

| Wind Characteristics Pavesi - 63 -


This is further calibrated in terms of wind velocity, which can be directly sensed and recorded.
Although these anemometers can sustain a variety of harsh environments, they have some limitations.
It accelerates quickly with the wind but retards slowly as wind ceases. Due to this slow response, cup
anemometers do not give reliable measurement in wind gusts. As the drag force is proportional to the
density, any changes in the air density will affect the accuracy of metered velocity. Nevertheless of
these limitations, cup anemometers are widely used for measuring wind velocity in meteorological as
well as wind energy applications.
The bridled anemometers are similar to the cup anemometers, but have more number of cups
(commonly 32). The rotation of the central axis to which the cups are attached is checked by a spring.
As the cups tend to rotate due to wind, tension builds up in the spring. Intensity of this tension is
translated into the wind speed.

4.5.2.2 Propeller anemometer

A propeller type anemometer (aero-vane) consists of a four bladed propeller. The blades are made
with light weight materials like aluminium or carbon fibre thermo plastic (CFT). These devices work
predominantly on lift force. With an air flow parallel to its axis, the propeller blades experiences a lift
force, which turns the propeller at a speed proportional to the wind velocity. For measuring the
horizontal and vertical components of wind, three propellers may be fixed on a common mast. The
response to any deviation in the direction of wind from the
propeller axis follows the cosine law. This means that the
velocity perpendicular to the propeller axis would be sensed as
zero.

4.5.2.3 Pressure plate anemometer

The first anemometer of any kind is the pressure plate


anemometer. It was invented by Leon Battista Alberti as early as
in 1450. This is further refined by Robert Hooke (1664) and
Rojer Pickering (1744). It basically consists of a swinging plate
held at the end of a horizontal arm. This is attached to a vertical
shaft around which the arm can rotate freely (Figure 4.5.3). A
Figure 4.5.3 Principle of pressure wind vane directs the plate always perpendicular to the wind
plate anemometer flow. As the drag coefficient of a flat plate can be taken as unity,
referring Eq. (5.5.2), the pressure exerted on the plate p by the
wind is given by
1
p  u 2 (5.5.2)
2
This pressure makes the plate to swing inward. As the
distance through which the plate swings depends on the wind
strength, it can be directly calibrated in terms of the wind
velocity. Pressure plate anemometers are suited for measuring
Figure 4.5.4 Principle of pressure gusty winds.
tube anemometer

| Wind Characteristics Pavesi - 64 -


4.5.2.4 Pressure tube anemometers

Another type of anemometer which uses the wind pressure to measure the velocity is the pressure
tube anemometer. This works on the principle that, the wind flow passing through the tube creates
pressure where as the flow across a tube results in suction. Consider two tubes as shown in the Figure
4.5.4. The pressure in the tube parallel to the wind is the sum of atmospheric pressure and the wind
pressure. Thus
u2
p1  p  C1 (5.5.3)
2
Similarly in the tube perpendicular to the wind, the pressure is
u2
p1  p  C2 (5.5.4)
2
where p is the atmospheric pressure and C1 and C2 are coefficients. Subtracting p2 from p1 and
solving for u, we get

 2  p1  p 2  
0.5

u  (5.5.5)
   C1  C2  
Thus, by measuring the difference in pressure inside the two tubes, the wind velocity can be
estimated. Values of C1 and C2 are available with the instrument. The pressure is measured using
standard manometers or pressure transducers. The major advantage of pressure tube anemometer is that
it does not have any moving parts. This anemometer has limited application in the open field
measurements as the presence of dust, moisture and insects can affect its accuracy.

4.5.2.5 Sonic anemometer

Sonic anemometers measure the wind velocity by


sensing the changes in the speed of sound in air. It has
three arms, mounted perpendicular to each other, as shown
in Figure 4.5.5. Transducers fitted at the tips of each arm
emit acoustic signals which travel up and down through the
air. Speed of sound in moving air is different from that
through still air. Let uS be the velocity of sound in still air
and u is the wind velocity. If both the sound and wind are
moving in the same direction, then the resultant speed of
sound waves (u1) is
u1  u s  u (5.5.6)
Similarly, if the propagation of the sound waves is
opposite to the wind direction, then the resultant velocity of
sound (u2) is

Figure 4.5.5 Sonic anemometer

| Wind Characteristics Pavesi - 65 -


u 2  us  u (5.5.7)
From Eq. (5.5.6)and Eq. (5.5.7)we get
u1  u 2
u (5.5.8)
2
Thus, by measuring the speed of sound waves between the transducer tips during its upward and
downward travel, the wind velocity can be estimated. Sonic anemometers also do not have any moving
parts. They are reliable and accurate for measuring wind velocity in the range of 0 to 65 m/s. However,
they are costlier than the other types of anemometers.
Some other anemometers are bridled anemometers and hot wire anemometers. Hot wire
anemometers utilize the cooling effect of wind for measuring its velocity. The rate at which a hot wire,
maintained at a temperature, is cooled depends on the velocity of air moving over it. But they are not
common in wind energy measurements.
Apart from the sensors discussed above, the complete anemometer assembly includes transducers,
data loggers and processing units. For example, the cup and propeller anemometers generally drive a
mini generator to convert the rotational speed in to electrical signals. Some designs are fitted with dc
generators whereas some recent models employ permanent magnet ac generators. The output voltage of
the generator is directly proportional to the wind speed. With an analog to digital (A/D) converter, the
analog voltage can be converted to digital form which can further be processed and stored by a suitable
data processing system.
The data processing unit in modern wind measurement systems includes an electronic chip on a
computer which receives the instantaneous data and averages it over a specific time period. The
standard time interval is 10 minutes. The instantaneous turbulences, which are not of interest in wind
energy conversion, smoothen out within this 10 minutes interval. At the same time, it can take care of
the dynamic changes that are significant for wind energy systems.
For reliable operation, periodic calibration is essential for all anemometers. This is done under
ideal conditions against a bench mark anemometer which is considered as the standard. Even with
proper calibration, it is observed that some errors may creep in to the measurements. One possible
cause for error is the tower shadow. The nearby structures, or even the anemometer tower itself, may
shade the instrument resulting in faulty readings. To minimize the risk of tower shadow, cylindrical
poles supported by guy wires are preferred than lattice towers. Another problem may be missing data
due to functional errors of the recording mechanism.
The quality of data measured using an anemometer will depend on the characteristics of the
instrument such as the accuracy, resolution, sensitivity, error, response speed, repeatability and
reliability. Accuracy of an anemometer is the degree to which the wind is being measured by the
instrument. For example, a typical cup type anemometer can have an accuracy of ± 0.3 m/s. Resolution
is the smallest change in wind speed that can be detected by the anemometer and sensitivity is the ratio
between the output and input signals. Error is the deviation of indicated velocity from the actual
velocity. The response speed tells us how quickly the changes in wind speed can be detected by the
anemometer. Repeatability indicates the closeness of readings given by the anemometer while repeated
measurements are made under identical conditions. Reliability shows the probability that the
anemometer works successfully within its given range of wind speeds. The anemometers used for wind
measurements should be periodically checked for these properties.

| Wind Characteristics Pavesi - 66 -


4.5.2.6 Wind direction

Direction of wind is an important factor in the sitting of a wind energy conversion system. If we
receive the major share of energy available in the wind from a certain direction, it is important to avoid
any obstructions to the wind flow from this side. Wind vanes were used to identify the wind direction
in earlier day’s anemometers. However, most of the anemometers used today have provisions to record
the direction of wind along with its velocity. A typical arrangement to measure the wind direction used
with cup anemometers is shown in Figure 4.5.6. Information
on the velocity and direction of wind, in a combined form, can
be presented in the wind roses. The wind rose is a chart which
indicates the distribution of wind in different directions. The
chart is divided into 8, 12 or even 16 equally spaced sectors
representing different directions. Three types of information
can be presented in a wind rose.
The percentage of time for which we receive wind from
a particular direction. This can show us the direction from
which we get most of our wind.
The product of this percentage and the average wind
velocity in this direction. This tells us the average strength of
the wind spectra.
The product of time percentage and cube of the wind
velocity. This helps us in identifying the energy available Figure 4.5.6 Vanes to measure wind
from different directions. Typical wind roses for a location are direction
shown in Error! Reference source not found.
…………………………..

| Wind Characteristics Pavesi - 67 -


4.6 Analysis of wind data
For estimating the wind energy potential of a site, the wind data collected from the location
should be properly analyzed and interpreted. Long term wind data from the meteorological stations
near to the candidate site can be used for making preliminary estimates. This data, which may be
available for long periods, should be carefully extrapolated to represent the wind profile at the potential
site. After this preliminary investigation, field measurements are generally made at the prospective
location for shorter periods. One year wind data recorded at the site is sufficient to represent the long
term variations in the wind profile within an accuracy level of 10 per cent.
Modern wind measurement systems give us the mean wind speed at the site, averaged over a pre-
fixed time period. Ten minutes average is very common as most of the standard wind analysis
softwares are tuned to handle data over ten minutes. This short term wind data are further grouped and
analyzed with the help of models and softwares to make precise estimates on the energy available in
the wind. The data are grouped over time spans in which we are interested in. For example, if we want
to estimate the energy available at different hours, then the data should be grouped in an hourly basis.
The data may also be categorized on daily, monthly or yearly basis.

4.6.1 Average wind speed


The speed of the wind is continuously changing, making it desirable to describe the wind by
statistical methods. Here we will examine a few of the basic concepts of probability and statistics,
leaving a more detailed treatment to the books written on the subject. One statistical quantity which we
have mentioned is the average or arithmetic mean. If we have a set of numbers ui, such as a set of
measured wind speeds, the mean of the set is defined as
1 n
u  ui
n i 1
(5.6.1)

The sample size or the number of measured values is n. Another quantity seen occasionally in the
literature is the median. If n is odd, the median is the middle number after all the numbers have been
arranged in order of size. As many numbers lie below the median as above it. If n is even the median is
halfway between the two middle numbers when we rank the numbers.
However, for wind power calculations, averaging the velocity using Eq. (5.6.1) is often
misleading and the wind power potential is underestimated.
For wind energy calculations, the velocity should be weighed for its power content while
computing the average. Thus, the average wind velocity is given by
1
1 n  3
u m    u i3  (5.6.2)
 n i 1 
This shows that due to the cubic velocity power relationship, the weighted average expressed in
Eq. (5.6.2) should be used in wind energy analysis.
In addition to the mean, we are interested in the variability of the set of numbers. We want to find
the discrepancy or deviation of each number from the mean and then find some sort of average of these
deviations. The mean of the deviations ui −u is zero, which does not tell us much. We therefore square
each deviation to get all positive quantities. The variance σ2 of the data is then defined as

| Wind Characteristics Pavesi - 68 -


1 n
2    ui  u 
2
(5.6.3)
n  1 i 1
The term n - 1 is used rather than n for theoretical reasons we shall not discuss here [Burr, I. W.
(1974)]. The standard deviation σ is then defined as the square root of the variance.
standard deviation = √variance (5.6.4)
Lower values of  indicate the uniformity of the data set.
Wind speeds are normally measured in integer values, so that each integer value is observed
many times during a year of observations. The numbers of observations of a specific wind speed ui will
be defined as mi. The mean is then
1 w
u  mi u i
n i 1
(5.6.5)

where w is the number of different values of wind speed observed and n is still the total number of
observations. It can be shown [2] that the variance is given by

1 w  
2
1 w
2   i i 2
  i i  
n  i 1
m u m u (5.6.6)
n  1  i 1 
Both the mean and the standard deviation will vary from one period to another or from one
location to another. It may be of interest to arrange these values in rank order from smallest to largest.
We can then immediately pick out the smallest, the median, and the largest value. The terms smallest
and largest are not used much in statistics because of the possibility that one value may be widely
separated from the rest. The fact that the highest recorded surface wind speed is known is not very
helpful in estimating peak speeds at other sites, because of the large gap between this speed and the
peak speed at the next site in the rank order. The usual practice is to talk about percentiles, where the
90 percentile mean wind would refer to that mean wind speed which is exceeded by only 10 % of the
measured means. Likewise, if we had 100 recording stations, the 90 percentile standard deviation
would be the standard deviation of station number 90 when numbered in ascending rank order from the
station with the smallest standard deviation. Only 10 stations would have a larger standard deviation (or
more variable winds) than the 90 percentile value. This practice of using percentiles allows us to
consider cases away from the median without being too concerned about an occasional extreme value.
We shall now define the probability p of the discrete wind speed ui being observed as
mi
p  ui   (5.6.7)
n
With this definition, the sum of all probabilities will be unity.
w

 pu   1
i 1
i (5.6.8)

We shall also define a cumulative distribution function F(ui) as the probability that a measured
wind speed will be less than or equal to ui.

| Wind Characteristics Pavesi - 69 -


Figure 4.6.1 General relationship between (a) a density function f(u) and (b) a distribution function
F(u)
j
F  ui    p  ui  (5.6.9)
i 1

The cumulative distribution function has the properties


F(−∞) = 0, F(∞) = 1 (5.6.10)
We also occasionally need a probability that the wind speed is between certain values or above a
certain value. We shall name this probability P(ua ≤ u ≤ ub) where ub may be a very large number. It is
defined as
b
P  ua  u  u b    p  ui  (5.6.11)
i a

It is convenient for a number of theoretical reasons to model the wind speed frequency curve by a
continuous mathematical function rather than a table of discrete values. When we do this, the
probability values p(ui) become a density function f(u). The density function f(u) represents the
probability that the wind speed is in a 1 m/s interval centred on u. The discrete probabilities p(ui) have
the same meaning if they were computed from data collected at 1 m/s intervals. The area under the
density function is unity, which is shown by the integral equivalent of Eq. (5.6.8).

 f  u  du  1
0
(5.6.12)

The cumulative distribution function F(u) is given by


u
F  u    f  x  dx (5.6.13)
0

The variable x inside the integral is just a dummy variable representing wind speed for the
purpose of integration. Both of the above integrations start at zero because the wind speed cannot be
negative.
When the wind speed is considered as a continuous random variable, the cumulative distribution

| Wind Characteristics Pavesi - 70 -


function has the properties F(0) = 0 and F(∞) = 1. The quantity F(0) will not necessarily be zero in the
discrete case because there will normally be some zero wind speeds measured which are included in
F(0). In the continuous case, however, F(0) is the integral of Eq. (5.6.13) with integration limits both
starting and ending at zero. Since f(u) is a well behaved function at u = 0, the integration has to yield a
result of zero. This is a minor technical point which should not cause any difficulties later. We will
sometimes need the inverse of Eq. (5.6.13) for computational purposes. This is given by
dF  u 
f u  (5.6.14)
du
The general relationship between f(u) and F(u) is sketched in Figure 4.6.1. F(u) starts at 0,
changes most rapidly at the peak of f(u), and approaches 1 asymptotically. The mean value of the
density function f(u) is given by

u   uf  u  du (5.6.15)
0

The variance is given by



2    u  u  f  u  du
2
(5.6.16)
0

These equations are used to compute theoretical values of mean and variance for a wide variety of
statistical functions that are used in various applications.

4.6.2 Weibull Statistics


If we join the midpoints of the frequency and cumulative histograms in Figure 4.6.1, we get
smooth curves with a well defined pattern. This shows that it is logical to represent the wind velocity
distributions by standard statistical functions. There are several density functions which can be used to
describe the wind speed frequency curve. The two most common are the Weibull and the Rayleigh
functions. For the statistically inclined reader, the Weibull is a special case of the Pearson Type III or
generalized gamma distribution, while the Rayleigh [or chi with two degrees of freedom (chi-2)]
distribution is a subset of the Weibull. The Weibull is a two parameter distribution while the Rayleigh
has only one parameter. This makes the Weibull somewhat more versatile and the Rayleigh somewhat
simpler to use. We shall present the Weibull distribution first. The wind speed u is distributed as the
Weibull distribution if its probability density function is

ku
k 1
  u k 
f u    exp      (k>0, u>0, c>1) (5.6.17)
cc   c  
This is a two parameter distribution where c and k are the scale parameter and the shape
parameter, respectively. Curves of f(u) are given in Figure 4.6.2, for the scale parameter c = 1. It can be
seen that the Weibull density function gets relatively more narrow and more peaked as k gets larger.
The peak also moves in the direction of higher wind speeds as k increases. A comparison of Figure
4.6.1 shows that the Weibull has generally the right shape to fit wind speed frequency curves, at least
for these two locations. Actually, data collected at many locations around the world can be reasonably
well described by the Weibull density function if the time period is not too short. Periods of an hour or

| Wind Characteristics Pavesi - 71 -


two or even a day or two

.
may have wind data 1.2 k=3.00
which are not well fitted k=2.50
by a Weibull or any 1.0
other statistical function, k=2.00
0.8

f(u)
but for periods of several k=1.50
weeks to a year or more,
the Weibull usually fits 0.6
k=1.25
the observed data
reasonably well. It may 0.4
k=1.00
have been noticed in
0.2
Figure 4.6.2 that the
wind speed only varies 0.0
between 0 and 2.5 m/s, a
range with little interest 0 0.5 1 1.5 2 2.5
u
from a wind power
viewpoint. This is not Figure 4.6.2 Weibull density function f(u) for scale parameter c = 1.
really a problem because the scale parameter c can scale the curves to fit different wind speed regimes.
If c is different from unity, the values of the vertical axis have to divided by c, as seen by Eq.
(5.6.17). Since one of the properties of a probability density function is that the area under the curve
has to be unity, then as the curve is compressed vertically, it has to expand horizontally.
For k greater than unity, f(u) becomes zero at zero wind speed. The Weibull density function thus
cannot fit a wind speed frequency curve at zero speed because the frequency of calms is always greater
than zero. This is not a serious problem because a wind turbine’s output would be zero below some cut-
in speed anyway. What is needed is a curve which will fit the observed data above some minimum
speed. The Weibull density function is a suitable curve for this task.
A possible problem in fitting data is that the Weibull density function is defined for all values of u
for 0 ≤ u≤∞ whereas the actual number of observations will be zero above some maximum wind speed.
Fitting a nonzero function to zero data can be difficult. Again, this is not normally a serious problem
because f(u) goes to zero for all practical purposes for u/c greater than 2 or 3, depending on the value of
k. Both ends of the curve have to receive special attention because of these possible problems, as will
be seen later. The mean wind speed u computed from Eq. (5.6.15) is

u   uf  u  du
0
(5.6.18)

ku
k 1
  u k 
 u   exp      du
0
cc   c  
If we make the change of variable
k
u
x   (5.6.19)
x
then the mean wind speed can be written

| Wind Characteristics Pavesi - 72 -



u  c x1 k e x dx (5.6.20)
0

It is basically in the form the gamma function. Tables of values exist for the gamma function and
it is also implemented in the large mathematical software packages just like trigonometric and
exponential functions. The gamma function , Γ(y), is usually written in the form

  y    z y1e z dz (5.6.21)
0

Equations (5.6.20) and (5.6.21) have the same integrand if y = 1+1/k. Therefore the mean wind
speed is
 1
u  c 1   (5.6.22)
 k
Published tables that are available for the gamma function Γ(y) are only given for 1 ≤ y ≤ 2. If an
argument y lies outside this range, the recursive relation
Γ(y + 1) = yΓ(y) (5.6.23)
must be used. If y is an integer,
Γ(y + 1) = y! = y(y − 1)(y − 2) … (1) (5.6.24)
The factorial y! is implemented on the more powerful hand calculators. The argument y is not
restricted to an integer, so the quantity computed is actually Γ(y + 1). This may be the most convenient
way of calculating the gamma function in many situations. Normally, the wind data collected at a site
will be used to directly calculate the mean speed u . We then want to find c and k from the data. A
good estimate for c can
be obtained quickly 1.2
from Eq. (5.6.22) by
considering the function 1.0
c/u

c/ u as a function of k
which is given inFigure 0.8
4.6.3. For values of k
below unity, the ratio c/ 0.6
u decreases rapidly.
For k above 1.5 and less 0.4
than 3 or 4, however,
the ratio c/ u is 0.2
essentially a constant,
with a value of about 0.0
1.12. This means that 0.0 0.5 1.0 1.5 2.0 2.5 3.0
the scale parameter is k
directly proportional to
the mean wind speed for Figure 4.6.3 Weibull scale parameter c divided by mean wind speed u
this range of k. versus Weibull shape parameter k

| Wind Characteristics Pavesi - 73 -


c = 1.12 u (1.5 ≤ k ≤ 3.0) (5.6.25)
Most good wind regimes will have the shape parameter k in this range, so this estimate of c in
terms of u will have wide application. It can be shown by substitution that the Weibull distribution
function F(u) which satisfies Eq. (5.6.14), and also meets the other requirements of a distribution
function, i.e. F(0) = 0 and F(∞) = 1, is
  u k 
F  u   1  exp      (5.6.26)
  c  
The variance of the Weibull density function can be shown to be
  2  1 
2  c 2  1     2 1   
  k  k 
  2 
  1  k  
(5.6.27)
 u     1
2


 2 1   
1
  k  

The probability of the wind speed u being equal to or greater than ua is



P  u  u a    f  u  du
ua

  u k  (5.6.28)
 exp     
  c  
The probability of the wind speed being within a 1 m/s interval centred on the wind speed ua is
u a  0.5
P  u a  0.5  u  u a  0.5   f  u  du
u a  0.5

  u  0.5 k    u a  0.5 k 
 exp    a    exp     (5.6.29)
  c     c  
f  u a  u  f  u a 
We shall see that the power in the wind passing through an area A perpendicular to the wind is
given by
1
Pw  Au 3
2
 W (5.6.30)

The average power in the wind is then


w
1
P w  a  p  u i u13
2 i 1
 W (5.6.31)

We may think of this value as the true or actual power in the wind if the probabilities p(ui) are

| Wind Characteristics Pavesi - 74 -


determined from the actual wind speed data. If we model the actual wind data by a probability density
function f(u), then the average power in the wind is
1 
P w  A  u 3f  u  du
2 0
 W (5.6.32)

It can be shown [Justus, C. G. (1978)] that when f(u) is the Weibull density function, the average
power is
3  3
Au  1  
 k
Pw 3 W (5.6.33)
  1 
2  1   
  k 
If the Weibull density function fits the actual wind data exactly, then the power in the wind
predicted by Eq. (5.6.33) will be the same as that predicted by Eq. (5.6.31). The greater the difference
between the values obtained from these two equations, the poorer is the fit of the Weibull density
function to the actual data. Actually, wind speeds outside some range are of little use to a practical
wind turbine. There is inadequate power to spin the turbine below perhaps 5 or 6 m/s and the turbine
may reach its rated power at 12 m/s. Excess wind power is spilled or wasted above this speed so the
turbine output power can be maintained at a constant value. Therefore, the quality of fit between the
actual data and the Weibull model is more important within this range than over all wind speeds. We
shall consider some numerical examples of these fits later in the chapter.
The function u3f(u) starts at zero at u = 0, reaches a peak value at some wind speed ume, and
finally returns to zero at large values of u. The yearly energy production at wind speed ui is the power
in the wind times the fraction of time that power is observed times the number of hours in the year. The
wind speed ume is the speed which produces more energy (the product of power and time) than any
other wind speed. Therefore, the maximum energy obtained from any one wind speed is
1
Wmax  Au 3me f  u m  [W] (5.6.34)
2
The turbine should be designed so this wind speed with maximum energy content is included in
its best operating wind speed range. Some applications will even require the turbine to be designed
with a rated wind speed equal to this maximum energy wind speed. We therefore want to find the wind
speed ume. This can be found by multiplying Eq. (5.6.17) by u3, setting the derivative equal to zero,
and solving for u. After a moderate amount of algebra the result can be shown to be
1
 k  2 k
u me  c   m s (5.6.35)
 k 
We see that ume is greater than c so it will therefore be greater than the mean speed u . If the
mean speed is 6 m/s, then ume will typically be about 8 or 9 m/s. We see that a number of interesting
results can be obtained by modelling the wind speed histogram by a Weibull density function.

| Wind Characteristics Pavesi - 75 -


4.6.2.1 Determining the Weibull Parameters

There are several methods available for determining the Weibull parameters c and k [Justus, C. G.
(1978)]. If the mean and variance of the wind speed are known, then Eqs. (5.6.22) and (5.6.27) can be
solved for c and k directly. At first glance, this would seem impossible because k is buried in the
argument of a gamma function. However, Justus, C. G. (1978) has determined that an acceptable
approximation for k from Eq. (5.6.27) is
1.086

k   (5.6.36)
u
This is a reasonably good approximation over the range 1 ≤k≤ 10. Once k has been determined,
Eq. (5.6.22) can be solved for c.
u
c (5.6.37)
 1
 1  
 k
The variance of a histogram of wind speeds is not difficult to find from Eq. (5.6.6), so this
method yields the parameters c and k rather easily.
Similarly, c can be approximated as
2u
c (5.6.38)

More accurately, c can be found using the expression [1]


u k 2.6674
c (5.6.39)
0.184  0.816 k 2.7385
The method can even be used when the variance is not known, by simply estimating k. Justus, C.
G. (1978) examined the wind speed distributions at 140 sites across the continental United States
measured at heights near 10 m, and found that k appears to be proportional to the square root of the
mean wind speed.

k  d1 u (5.6.40)
The proportionality constant d1 is a site specific constant with an average value of 0.94 when the
mean wind speed u is given in meters per second. The constant d1 is between 0.73 and 1.05 for 80 %
of the sites. The average value of d1 is normally adequate for wind power calculations, but if more
accuracy is desired, several months of wind data can be collected and analyzed in more detail to
compute c and k. These values of k can be plotted versus u on log-log paper, a line drawn through
the points, and d1 determined from the slope of the line.
Another method of determining c and k which lends itself to computer analysis, is the least
squares approximation to a straight line. In the graphical method, we transform the cumulative
distribution function in to a linear form, adopting logarithmic scales [5,18]. That is, we perform the

| Wind Characteristics Pavesi - 76 -


necessary
mathematical
operations on Eq.
(5.6.17) to linearize it
and then determine c
and k to minimize the
least squared error
between the linearized
ideal curve and the
actual data points of
p(ui).
The first step of
linearization is to
integrate Eq. (5.6.14).
This yields the
distribution function
F(u) which is given by
Eq. (5.6.26).
The expression
for the cumulative
distribution of wind
velocity can be Figure 4.6.4 Cumulative distribution generated using different methods
rewritten as
  u k 
1  F  u   exp      (5.6.41)
  c  
F(u) is more nearly describable by a straight line than f(u), but is still quite nonlinear. We note
that F(u) contains an exponential and that, in general, exponentials are linearized by taking the
logarithm. In this case, because the exponent is itself raised to a power, we must take logarithms twice.

 
ln  ln 1  F  u   k ln  u   k ln  c  (5.6.42)

This is in the form of an equation of a straight line


y = ax + b (5.6.43)
where x and y are variables, a is the slope, and b is the intercept of the line on the y axis. In
particular,
y = ln[-ln(1 - F(u)]
a=k
(5.6.44)
x = ln u
b = -k ln c
Data will be expressed in the form of pairs of values of ui and F(ui). For each wind speed ui there
is a corresponding value of the cumulative distribution function F(ui). When given values for u = ui and

| Wind Characteristics Pavesi - 77 -


F(u) = F(ui) we can find values for x = xi and y = yi in Eq. 55. Being actual data, these pairs of values
do not fall exactly on a straight line, of course. The idea is to determine the values of a and b in Eq.
(5.6.43) such that a straight line drawn through these points has the best possible fit. It can be shown
that the proper values for a and b are
w w

w  x i  yi
x y  i i
i 1
w
i 1

a i 1
2
  w

  xi 
x i2   i 1 
w


i 1 w
(5.6.45)

x  
w

i  x yi  y
 i 1

x 
w 2
i x
i 1

b  yi  ax i
1 w a w (5.6.46)
  i w
w i 1
y 
i 1
xi

In these equations x and y are the mean values of xi and yi, and w is the total number of pairs
of values available. The final results for the Weibull parameters are
ka
 b (5.6.47)
c  exp   
 k
One of the implied assumptions of the above process is that each pair of data points is equally
likely to occur and therefore would have the same weight in determining the equation of the line. For
typical wind data, this means that one reading per year at 20 m/s has the same weight as 100 readings
per year at 5 m/s. To remedy this situation and assure that we have the best possible fit through the
range of most common wind speeds, it is possible to redefine a weighted coefficient a in place of Eq.
(5.6.45) as

 p u x  
w
2
i i  x yi  y
a i 1
(5.6.48)
 p u x 
w 2
2
i i x
i 1

This equation effectively multiplies each xi and each yi by the probability of that xi and that yi
occurring. It usually gives a better fit than the unweighted a of Eq. (5.6.45). Eqs. (5.6.45), (5.6.46), and
(5.6.48) can be evaluated conveniently on a programmable hand held calculator. Some of the more
expensive versions contain a built-in linear regression function so Eqs. (5.6.45) and (5.6.46) are
handled internally. All that needs to be entered are the pairs of data points. This linear regression

| Wind Characteristics Pavesi - 78 -


function can be combined with the programming capability to evaluate Eq. (5.6.48) more conveniently
than by separately entering each of the repeating data points mi times.
We can see that, there are differences in k and c estimated, following the graphical and standard
deviation methods. The cumulative distributions generated using these methods are compared with the
field measurements (represented by dotted lines) in Figure 4.6.4. Results of the graphical method are
found closer to the field observations. The standard deviation method also could predict the wind
distribution within an acceptable level of accuracy.
Another method for estimating k and c is the First and Second Order Moment Method. The nth
moment Mn of Weibull distribution is given by
 n
M n  c n  1   (5.6.49)
 k
If M1 and M2 are the first and second moments, with Eq.(5.6.49), c can be solved as
 1
 1  
c 2 
M k
(5.6.50)
M1  2 
 1  
 k
Similarly,
 1
 1  
 
M2 k
(5.6.51)
2
M1  2
 2 1  
 k
In this method, M1 and M2 are calculated from the given wind data. Further, k and c are estimated
by solving Eq. (5.6.50) and Eq.(5.6.51).
Maximum likelihood method
In the maximum likelihood method, the shape and scale factors are given as [3, 19]
1
 n k n

  u i ln  u i   u ik 
k   i 1 n  i 1  (5.6.52)
 n 
 
u ik
i 1


1
1 n  k
c    u ik  (5.6.53)
 n i 1 
Energy pattern factor (EPF) is the ratio between the total power available in the wind and the
power corresponding to the cube of the mean wind speed [14]. Thus

| Wind Characteristics Pavesi - 79 -


1 n 3
u
n i 1 i
E PF  3
(5.6.54)
1 n 
 n  ui 
 i 1 
Once the energy pattern factor for a regime is found from the wind data, an approximate solution
for k is
k = 3.957 EPF - 0.898 (5.6.55)
Various methods used for estimating k and c are discussed in detail in [19].

4.6.3 Rayleigh and Normal Distributions


We now return to a discussion of the other popular probability density function in wind power
studies, the Rayleigh or chi-2. The reliability of Weibull distribution in wind regime analysis depends
on the accuracy in estimating k and c. For the precise calculation of k and c, adequate wind data,
collected over shorter time intervals are essential. In many cases, such information may not be readily
available. The existing data may be in the form of the mean wind velocity over a given time period (for
example daily, monthly or yearly mean wind velocity). Under such situations, a simplified case of the
Weibull model can be derived, approximating k as 2. This is known as the Rayleigh distribution.
The Rayleigh probability density function is given by

u    u 2 
f  u   2 exp      (5.6.56)
2u  4  u  
The Rayleigh cumulative distribution function is
   u 2 
F  u   1  exp      (5.6.57)
 4  u  
The probability that the wind speed u is greater than or equal to ua is just
   u 2 
P  u  u a   1  F  u   exp      (5.6.58)
 4  u  
The variance of this density function is
4  2
 2    1 u (5.6.59)
 
It may be noted that the variance is only a function of the mean wind speed. This means that one
important statistical parameter is completely described in terms of a second quantity, the mean wind
speed. Since the mean wind speed is always computed at any measurement site, all the statistics of the
Rayleigh density function used to describe that site are immediately available without massive amounts
of additional computation. As mentioned earlier, this makes the Rayleigh density function very easy to
use. The only question is the quality of results. It appears from some studies [Corotis, R. B. (1977),
Corotis, R. B., et al. (1978)], that the Rayleigh will yield acceptable results in most cases. The natural

| Wind Characteristics Pavesi - 80 -


variability in the wind from year to year tends to
limit the need for the greater sophistication of
the Weibull density function. It should be
emphasized that actual histograms of wind
speeds may be difficult to fit by any
mathematical function, especially if the period of
time is short. This is illustrated by Figure 4.6.5,
which shows actual data for a site and the
Weibull and Rayleigh models of the data.
The wind speed was automatically
measured about 245,000 times over a 29 day
period in July, 1980, between 7:00 and 8:00 p.m.
to form the histogram. A bimodal (two-humped)
characteristic is observed, since 3 m/s and 9 m/s
are observed more often than 5 m/s. This is
evidently due to the wind speed being high for a

Figure 4.6.6 Actual wind data, and Weibull and


Rayleigh density functions for Tuttle Creek, 7:00–
8:00 p.m., July 1980, 50 m.
number of days and relatively low the other days,
with few days actually having average wind speeds.
The Weibull is seen to be higher than the
Rayleigh between 5 and 12 m/s and lower outside
this range. Both functions are much higher than the
actual data above 12 m/s. The actual wind power
density as computed from Eq. (5.6.31) for standard
conditions is P w A  362 W/m2. The power density
computed from the Weibull model, Eq. (5.6.33), is
467 W/m2, while the power density computed from
the Rayleigh density function by a process similar to
Eq. (5.6.31) is 565 W/m2. The Weibull is 18 % high
while the Rayleigh is 43 % high.
Figure 4.6.5 Actual wind data, and Weibull If we examine just the wind speed range of 5-
and Rayleigh density functions for Tuttle Creek, 12 m/s, we get an entirely different picture. The
7:00–8:00 a.m., August 1980, 10 m. actual wind power density in this range is

| Wind Characteristics Pavesi - 81 -


P w A  362 W/m2, while the Weibull predicts a density of 308 W/m2, 15 % low, and the Rayleigh
predicts 262 W/m2, 28 % low. This shows that neither model is perfect and that results from such
models need to be used with caution. However, the Weibull prediction is within 20 % of the actual
value for either wind speed range, which is not bad for a data set that is so difficult to mathematically
describe.
Another example of the ability of the Weibull and Rayleigh density functions to fit actual data is
shown in Figure 4.6.6. The mean speed is 4.66 m/s as compared with 7.75 m/s in Figure 4.6.6 and k is
1.61 as compared with 2.56. This decrease in k to a value below 2 causes the Weibull density function
to be below the Rayleigh function over a central range of wind speeds, in this case 2–8 m/s. The actual
data are concentrated between 2 and 4 m/s and neither function is able to follow this wide variation.
The actual wind power density in this case is 168 W/m2, while the Weibull prediction is 161 W/m2, 4
% low, and the Rayleigh prediction is 124 W/m2, 26 % low. Considering only the 5–12 m/s range, the
actual power density is 89 W/m2, the Weibull prediction is 133 W/m2, 49 % high, and the Rayleigh
prediction is 109 W/m2, 22 % high.
We see from these two figures that it is difficult to make broad generalizations about the ability of
the Weibull and Rayleigh density functions to fit actual data. Either one may be either high or low in a
particular range. The final test or proof of the usefulness of these functions will be in their ability to
predict the power output of actual wind turbines. In the meantime it appears that either function may
yield acceptable results, with the Weibull being more accurate and the Rayleigh easier to use.
Although the actual wind speed distribution can be described by either a Weibull or a Rayleigh
density function, there are other quantities which are better described by a normal distribution. The
distribution of monthly or yearly mean speeds is likely to be normally distributed around a long-term
mean wind speed, for example. The normal curve is certainly the best known and most widely used
distribution for a continuous random variable, so we shall mention a few of its properties.
The density function f(u) of a normal distribution is
 uu
  
2

exp   
1
f u  (5.6.60)
 2  22 
 

where u the mean and σ is the standard deviation. In this expression, the variable u is allowed to
vary from −∞ to +∞. It is physically impossible for a wind speed to be negative, of course, so we
cannot forget the reality of the observed quantity and follow the mathematical model past this point.
This will not usually present any difficulty in examining mean wind speeds.
The cumulative distribution function F(u) is given by

F  u    f  x  dx
u
(5.6.61)


This integral does not have a simple closed form solution, so tables of values are determined from
approximate integration methods. The variable in this table is usually defined as
uu
q

or (5.6.62)
u  u  q

| Wind Characteristics Pavesi - 82 -


Thus q is the number of standard deviations that u is away from u . A brief version of this table is
shown in Table 4.6.1. We see from the table, for example, that F(u) for a wind speed one standard
deviation below the mean is 0.159. This means there is a 15.9% probability that the mean speed for any
period of interest will be more than one standard deviation below the long term mean. Since the normal
density function is symmetrical, there is also a 15.9% probability that the mean speed for some period
will be more than one standard deviation above the long term mean. As we move further away from the
mean, the probability decreases rapidly. There is a 2.3% probability that the mean speed will be smaller
than a value two standard deviations below the long term mean, and only 0.1 % probability that it will
be smaller than a value three standard
deviations below the long term mean. This gives Table 4.6.1 Normal Cumulative Distribution
us a measure of the width of a normal Function
distribution. q F (u) q F (u)
It is also of interest to know the probability -3.090 0.001 +0.43 0.666
of a given data point being within a certain -2.326 0.01 +0.675 0.75
distance of the mean. A few of these probabilities -2.054 0.02 +0.842 0.8
are shown in Table 4.6.2. For example, the
-2.00 0.023 +1.0 0.841
fraction 0.6827 of all measured values fall within
-1.645 0.05 +1.281 0.9
one standard deviation of the mean if they are
-1.281 0.1 +1.645 0.95
normally distributed. Also, 95 % of all measured
values fall within 1.96 standard deviations of the -1.0 0.159 +2.0 0.977
mean. We can define the em 90 percent -0.842 0.2 +2.054 0.98
confidence interval as the range within 1.645 -0.675 0.25 +2.326 0.99
standard deviations of the mean. If the mean is -0.43 0.334 +3.090 0.999
10 m/s and the standard deviation happens to be 0 0.5
1 m/s, then the 90% confidence interval will be between 8.355 and
11.645 m/s. That is, 90% of individual values would be expected to Table 4.6.2 Probability of
lie in this interval. finding the variable within q
We might now ask ourselves how confident we are in the results standard deviations of u .
of this example. After all, only one year’s wind data were examined. q P(|u − u |≤qσ)
Perhaps we picked an unusual year with mean and standard deviation
far removed from their long term averages. We need to somehow 1.000 0.6827
specify the confidence we have in such a result. Justus, Mani and 1.645 0.9000
Mikhail (1970) examined long term wind data for 40 locations in the 1.960 0.9500
United States, including Alaska, Hawaii, and Wake Island. All sites 2.000 0.9545
had ten or more years of data from a fixed anemometer location and a 2.576 0.9900
long term mean wind speed of 5 m/s or greater. 3.000 0.9973
They found that monthly and yearly mean speeds are distributed very closely to a normal or
Gaussian distribution, as was mentioned earlier. The monthly means were distributed around the long
term measured monthly mean u m with an average standard deviation of 0.098 u m where u m is the
mean wind speed for a given month of the year, e.g. all the April average wind speeds are averaged
over the entire period of observation to get a long term average for that month. For a normal
distribution the 90 % confidence interval would be, using Table 4.6.2, u m 0.645(0.098) u m or the
interval between 0.84 u m and 1.16 u m . We can therefore say that we have 90% confidence that a
measured monthly mean speed will fall in the interval 0.84 u m to 1.16 u m . If we say that each

| Wind Characteristics Pavesi - 83 -


measured monthly mean lies
in this interval, we will be
correct 90 times out of 100,
and wrong 10 times.
The above argument
applies on the average. That
is, it is valid for sites with an
average standard deviation.
We cannot be as confident of
sites with more variable
winds and hence higher
standard deviations. Since we
do not know the standard
deviation of the wind speed
Figure 4.6.7 Confidence intervals for two measured monthly means, at the candidate site, we have
u m1 and u m 2 . to allow for the possibility of
it being a larger number. According to Justus (1978), 90 % of all observed monthly standard deviations
were less than 0.145 u m . This is the 90 percentile level. The appropriate interval is now u m 
1.645(0.145) u m or the interval between 0.76 u m and 1.24 u m . That is, any single monthly mean speed
at this highly variable site will fall within the interval 0.76 u m and 1.24 u m with 90 % confidence. The
converse is also true. That is, if we designate the mean for one month as u m1 , the unknown long term
monthly mean u m will fall within the interval 0.76 u m1 and 1.24 u m1 with 90 % confidence. We shall
try to clarify this statement with Figure 4.6.7.
Suppose that we measure the average monthly wind speed for April during two successive years
and call these values u m1 and u m 2 . These are shown in Figure 4.6.7 along with the intervals 0.76 u m1
to 1.24 u m1 , 0.76 u m 2 to 1.24 u m 2 , and 0.76 u m to 1.24 u m . The confidence interval centred on u m
contains 90 % of measured means for individual months. The mean speed u m1 is below u m but its
confidence interval includes u m . The mean speed u m 2 is outside the confidence interval for u m and
the reverse is also true. If we say that the true mean u m is between 0.76 u m 2 and 1.24 u m 2 , we will be
wrong. But since only 10% of the individual monthly means are outside the confidence interval for u m
, we will be wrong only 10 % of the time. Therefore we can say with 90 % confidence that u m lies in
the range of 0.76 u mi to 1.24 u mi where u mi is some measured monthly mean.
The 90 % confidence interval for monthly means is rather wide. Knowing that the monthly mean
speed will be between 11 and 16 knots (5.7 and 8.2 m/s), as in the previous example, may not be of
much help in making economic decisions. We will see that the energy output of a given turbine will
increase by a factor of two as the monthly mean speed increases from 5.7 to 8.2 m/s. This means that a
turbine which is a good economic choice in a 8.2 m/s wind regime may not be a good choice in a 5.7
m/s regime. Therefore, one month of wind speed measurements at a candidate site will not be enough,
in many cases. The confidence interval becomes smaller as the number of measurements grows. The
wind speed at a given site will vary seasonally because of differences in large scale weather patterns
and also because of the local terrain at the site. Winds from one direction may experience an increase in
speed because of the shape of the hills nearby, and may experience a decrease from another direction. It

| Wind Characteristics Pavesi - 84 -


is therefore advisable to make measurements of speed over a full yearly cycle. We then get a yearly
mean speed u y . Each yearly mean speed u y will be approximately normally distributed around the
long term mean speed u . Justus (1978) found that the 90 % confidence interval for the yearly means
was between 0.9 u and 1.1 u for the median (50 percentile) standard deviation and between 0.85 u
and 1.15 u for the 90 percentile standard deviation. As expected, these are narrower confidence
intervals than were observed for the single monthly mean.
Our estimate of the long term mean speed by one year’s data can conceivably be improved by
comparing our one year mean speed to that of a nearby National Weather Service station. If the yearly
mean speed at the NWS station was higher than the long term mean speed there, the measured mean
speed at the candidate site can be adjusted upward by the same factor. This assumes that all the winds
within a geographical region of similar topography and a diameter up to a few hundred kilometres will
have similar year to year variations. If the long term mean speed at the NWS station is u , the mean for
one year is u a , and the mean for the same year at the candidate site is u b , then the corrected or
estimated long term mean speed u c at the candidate site is
u
uc  ub (5.6.63)
ua
Note that we do not know how to assign a confidence interval to this estimate. It is a single
number whose accuracy depends on both the accuracy of u and the correlation between u a and u b .
The accuracy of u should be reasonably good after 30 or more years of measurements, as is common
in many NWS stations. However, the assumed correlation between u a and u b may not be very good.
Justus (1978) found that the correlation was poor enough that the estimate of long term means was not
improved by using data from nearby stations. Equally good results would be obtained by applying the
90 % confidence interval approach as compared to using Eq. (5.6.63). The reason for this phenomenon
was not determined. One possibility is that the type of anemometer used by the National Weather
Service can easily get dirty and yield results that are low by 10 to 20 % until the next maintenance
period. One NWS station may have a few months of low readings one year while another NWS station
may have a few months of low readings the next year, due to the measuring equipment rather than the
wind. This would make a correction like Eq. (5.6.63) very difficult to use. If this is the problem, it can
be reduced by using an average of several NWS stations and, of course, by more frequent maintenance.
Other studies are necessary to clarify this situation.
The actual correlation between two sites can be defined in terms of a correlation coefficient r,
where

u  u 
w
ai  ua bi  ub
r i 1
(5.6.64)
u  u 
w 2 w 2
ai  ua bi  ub
i 1 i 1

In this expression, u a is the long term mean speed at site a, u b is the long term mean speed at
site b, u bi is the observed monthly or yearly mean at site b, and w is the number of months or years
being examined. We assume that the wind speed at site b is linearly related to the speed at site a. We

| Wind Characteristics Pavesi - 85 -


can then plot each mean speed u bi versus the corresponding u ai . We then find the best straight line
through this cluster of points by a least squares or linear regression process. The correlation coefficient
then describes how closely our data fits this straight line. Its value can range from r = +1 to r = −1. At r
=+1, the data falls exactly onto a straight line with positive slope, while at r = −1, the data falls exactly
onto a straight line with negative slope. At r = 0, the data cannot be approximated at all by a straight
line.
Table 4.6.3 Average yearly speeds corrected to 30 m (100 ft)
Year Dodge City Wichita Russell
1948 14.98 14.09 14.14a
1949 13.53 13.35 13.08b
1950 12.89 12.77 14.11
1951 13.83 13.47 13.07
1952 14.22 13.97 12.31
c
1953 15.57 14.45 13.37
1954 15.20 14.98 12.83
1955 14.83 15.22 14.89
1956 14.60 14.86 14.85
1957 14.71 13.60 15.38
1958 13.85 13.74 13.63
1959 14.91 15.21 14.88
1960 15.23 14.55 13.67
1961 12.37 13.29 12.45
1962 13.53 13.18 12.99
1963 14.90 12.74 15.20
1964 16.32 13.95 16.04
1965 15.17 13.46 14.71
1966 15.37 13.40 15.17
1967 15.26 13.78 15.20
1968 15.25 14.51 15.70d
1969 14.19 13.33 14.52e
1970 14.85 14.11 15.66
1971 15.04 13.97 15.01
1972 14.40 13.84 13.39
1973 15.35 13.87 14.41
a
Estimated from 50, 51, 52 data by method of ratios.
b
Estimated from 50, 51, 52 data by method of ratios.
c
Estimated from 50, 51, 52 data by method of ratios.
d
Estimated from 66, 67, 70, 71 data by method of ratios.
e
Estimated from 66, 67, 70, 71 data by method of ratios.

In examining the data in Table 4.6.3, one notices some clusters of low wind speed years and high
wind speed years. That is, it appears that a low annual mean wind may persist through the following
year, and likewise for a high annual mean wind. If this is really the case, we may select a site for a

| Wind Characteristics Pavesi - 86 -


wind turbine in spite of a poor wind year, and then find the performance of the wind turbine to be poor
in the first year after installation. This has economic implications because the cash flow requirements
will usually be most critical the first year after installation. The turbine may go ahead and deliver the
expected amount of energy over its lifetime, but this does not help the economic situation immediately
after installation. Justus (1978) examined this question also, for 40 sites in the United States, and
concluded that there is about a 60% probability that one low wind speed month or year will be followed
by a similar one. If wind speeds were totally random, then this probability would be 50 %, the same as
for two heads in a row when flipping a coin. The probability of an additional low speed month or year
following the first two is totally random. Based on these results, Justus concluded that clusters of bad
wind months or years are not highly probable. In fact, the variation of speed from one period to the
next is almost totally random, and therefore such clusters should not be of major concern.
We have seen that the 90% confidence interval for the long term annual mean speed at a site is
between 0.9 u and 1.1 u when only one year’s data is collected. If this is not adequate, then additional
data need to be collected. Corotis (1979) reports that the 90% confidence interval is between 0.93 u
and 1.07 u for two year’s data, and between 0.94 u and 1.06 u for three year’s data. The confidence
interval is basically inversely proportional to the square root of the time period, so additional years of
data reduce the confidence interval at a slower and slower rate.

4.6.4 Distribution of Extreme Winds


Two important wind speeds which affect turbine cost are the design wind speed which the rotor
can withstand in a parked rotor configuration, and the maximum operating wind speed. A typical wind
turbine may start producing power at 5 to 7 m/s (11 to 16 mi/h), reach rated power at 12 to 16 m/s (27
to 36 mi/h), and be shut down at a maximum operating speed of 20 to 25 m/s (45 to 56 mi/h).
The design wind speed is prescribed in a ANSI Standard, as modified by height and exposure.
This varies from 31 m/s (70 mi/h) to 49 m/s (110 mi/h) in various parts of the United States. It is not
uncommon for wind turbines to be designed for 55 m/s (125 mi/h) or more. This high design speed
may be necessary at mountainous or coastal sites, but may be unnecessary and uneconomical in several
Great Plains states where the highest wind speed ever measured by the National Weather Service is less
than 45 m/s (100 mi/h). Also of interest is the number of times during a year that the wind reaches the
maximum turbine operating wind speed so the turbine must be shut down. This affects the turbine
design and has an impact on the utility when the power output suddenly drops from rated to zero. For
example, if a wind farm had to be shut down during the morning load pickup, other generators on the
utility system must ramp up at a greater rate, perhaps causing operational problems.
Wind turbine designers and electric utility operators therefore need accurate models of extreme
winds, especially the number of times per year the maximum operating speed is exceeded. Variation of
extreme winds with height is also needed. The ANSI Standard gives the extreme wind that can be
expected once in 50 years, but daily and monthly extremes are also of interest.
The extreme wind data which are available in the United States are the monthly or yearly fastest
mile of wind at a standard anemometer height. The anemometer is obviously an averaging device. It
gives the average speed of the fastest mile, which will be smaller than the average speed of the fastest
1/10th mile, or any other fraction of a mile, because of the fluctuation of wind speed with time. But one
has to stop refining data at some point, and the fastest mile seems to have been that point for structural
studies.
The procedure is to fit a probability distribution function to the observed high wind speed data.

| Wind Characteristics Pavesi - 87 -


The Weibull distribution function of Eq. (5.6.17) could probably be used for this function, except that it
tends to zero somewhat too fast. It is convenient to define a new distribution function Fe(u) just for the
extreme winds. Weather related extreme events, such as floods or extreme winds [Gumbel, E. J.
(1958)], are usually described in terms of one of two Fisher-Tippett distributions, the Type I or Type II.
Thom [Thom, H. C. S. (1960), Thom, H. C. S. (1968)] used the Type II to describe extreme winds in
the United States. He prepared a series of maps based on annual extremes, showing the fastest mile for
recurrence intervals of 2, 50, and 100 years. Height corrections were made by applying the one-seventh
power law. National Weather Service data were used. The ANSI Standard is based on examination of a
longer period of record by Simiu E. (1979) and uses the Type I distribution. Only a single map is given,
for a recurrence interval of 50 years. Other recurrence intervals are obtained from this map by a
multiplying factor. Tables are given for the variation of wind speed with height. Careful study of a
large data set collected by Johnson and analyzed by Henry D. H. (1986) showed that the Type I
distribution is superior to the Type II, so the mathematical description for only the Type I will be
discussed here. The Fisher-Tippett Type I distribution has the form
Fe(u) = exp(−exp(−α(x − β))) (5.6.65)
where Fe(u) is the probability of the annual fastest mile of wind speed being less than u. The
parameters α and β are characteristics of the site that must be estimated from the observed data. If n
period extremes are available, the maximum likelihood estimate of α may be obtained by choosing an
initial guess and iterating
1
i 1  (5.6.66)
exp  i x j 
n

x
j1
j

x
 exp   x 
n

i j
j1

until convergence to some value  . The maximum likelihood estimate of β is


0.5772
x (5.6.67)

If Fe(u) = 0.5, then u is the median annual fastest mile. Half the years will have a faster annual
extreme mile and half the years will have a slower one. Statistically, the average time of recurrence of
speeds greater than this median value will be two years. Similar arguments can be made to develop a
general relationship for the mean recurrence interval Mr, which is
1
Mr  (5.6.68)
1  Fe  u 
which yields

1   1 
u ln  ln 1     (5.6.69)
   Mr 
A mean recurrence interval of 50 years would require Fe(u) = 0.98, for example.
If we have 20 years or more of data, then it is most appropriate to find the distribution of yearly

| Wind Characteristics Pavesi - 88 -


extremes. We use one value for each year and calculate α and β. If we have only a few years of data,
then we might want to find the distribution of monthly extremes. The procedure is the same, but we
now have 12 values for each year instead of one. If we have only a few months to a year of wind data,
then we can find the distribution of daily extremes. When modelling is conducted using one period
length and it is desired to express the results in terms of a larger period, the parameters of the
distribution must be altered. If there are n of the smaller periods within each larger unit, the distribution
function for the larger period is
FL = [Fs]n (5.6.70)
where Fs is the distribution function for the smaller period. The relationship between parameters is
then
ln n
 L  s and L  s  (5.6.71)
s
For example, the relationship between monthly and yearly extremes is
2.485
a   m and a  m  (5.6.72)
m
An example of values of α and β for daily extremes is shown in Table 4.6.4. These are calculated
from a large data set collected by Kansas State University for the period October, 1983 through
September, 1984 at 7 Kansas locations, using the daily fastest minute
Differences among sites are relatively small, as seen by the standard deviations. There is a
tendency for α to get smaller and β to get larger with height. Similar data for 5 Kansas National
Weather Service Stations for the same time period is shown in Table 4.6.4
The fastest mile is used rather than the fastest minute, which theoretically makes only a small
difference.
Differences among the NWS sites are also relatively small. However, the NWS sites have a
consistently larger α and smaller β than the KSU sites in Table 4.6.4. Both differences cause the
predicted extreme wind to be smaller at the NWS sites than at the KSU sites. In making a detailed
comparison of the Dodge City NWS and the Dodge City KSU sites, which are about 5 miles apart in
similar terrain, it appears that the difference was caused by the type of anemometer used. The KSU
study used the Maximum anemometer, a small, light, and inexpensive device, while the National
Table 4.6.4 α and β for seven Kansas sites, daily fastest minute in mi/h 50 m 30 m 10
m
α β α β α β
Tuttle Creek Lake 0.1305 23.47 0.1298 22.25 0.1414 20.76
Plainville 0.1410 26.04 0.1474 23.63 0.1660 19.57
Beaumont 0.1441 26.14 0.1522 24.71 0.1620 21.74
Lyndon 0.1582 23.00 0.1547 21.24 0.1648 18.21
Dodge City 0.1448 25.86 0.1483 23.92 0.1552 20.83
Lyons 0.1537 23.55 0.1594 21.75 0.1689 19.49
Oakley 0.1435 26.89 0.1457 24.51 0.1556 20.72
mean 0.1451 24.24 0.1482 23.14 0.1591 20.19
std. dev. 0.0089 2.16 0.0094 1.39 0.0094 1.17

| Wind Characteristics Pavesi - 89 -


Weather Service used the Electric Speed Indicator anemometer, a much larger and heavier device. The
10 m mean speed with the KSU equipment was 12.8 mi/h for the year, while the NWS mean was 14.1
mi/h, so the Maximum anemometer indicated a mean speed 9 % below the NWS mean. On the other
hand, the extreme winds measured by the Maximum anemometer averaged 18 % higher than the NWS
extremes. The Maximum anemometer indication for the extreme winds varied from about the same as
the NWS measurement to as much as 15 mi/h faster. It is evident that the type of measurement
equipment can make a significant difference in the results.
A longer period of time was then analyzed for two of the NWS Stations, using monthly extremes
rather than daily extremes. The results of this study are shown in Table 4.6.6.
The fastest mile observed at Dodge City for this period was 77 mi/h, while at Topeka it was 74
mi/h. The α values are much closer to the KSU values than to the NWS values for the one year study.
The β values are larger as required by Eq. (5.6.71). From these values for Dodge City, the extreme
wind expected once every 3.3 months (approximately 100 days) would be 49.23 mi/h, a value
surprisingly close to the prediction using the KSU data and daily extremes. As mentioned earlier, it is
important to have an estimate of the number of days during a year in which the wind speed should
equal or exceed a given value. This is done by inserting values of α, β, and the given wind speed in Eq.
(5.6.65), then using Eq. (5.6.66) to find Mr, and finally taking the reciprocal of the interval Mr to find
Table 4.6.5 α and β for five Kansas NWS Stations, daily fastest
mile.
α β Height in ft
Wichita 0.2051 17.01 25
Topeka 0.2068 14.89 20
Concordia 0.2149 17.45 33
Goodland 0.2272 18.20 20
Dodge City 0.2060 19.46 20
mean 0.2120 17.40
std. dev. 0.0094 1.68

Table 4.6.6 α and β for Topeka and Dodge City


using monthly fastest mile for 1949–1968.
α β
Topeka 0.1420 34.38
Dodge City 0.1397 41.94

Table 4.6.7 Days per year a given wind speed or greater is observed.
speed 50 m 30 m 10 m NWS Topeka Dodge City
30 128.4 110.8 69.1 39.3
35 69.1 57.8 33.0 14.6
40 35.3 28.8 15.3 5.3 4.4 8.8
45 17.5 14.0 7.0 1.9 2.4 5.8
50 8.6 6.8 3.2 0.7 1.2 3.3
55 4.2 3.2 1.4 0.2 0.6 1.8
60 2.0 1.6 0.6 0.1 0.3 0.9

| Wind Characteristics Pavesi - 90 -


the frequency. Values are given in Table 4.6.7 for the mean KSU parameters, the mean NWS
parameters from Table 4.6.5, and the mean NWS parameters for monthly extremes from Table 4.6.6,
all expressed in days per year for a wind speed in mi/h.
There is some scatter in the results, as has been previously discussed, but the results are still quite
acceptable for many purposes. For example, a wind turbine with a maximum operating speed of 45
mi/h and a height of 10 m or less would be expected to be shut down between 2 and 7 times per year. If
the turbine height were increased to 30 m, the number of shutdowns would approximately double, from
7 to 14. Increasing the maximum operating speed by 5 mi/h will cut the number of shutdowns in half.
These rules of thumb may be quite useful to system designers.
As mentioned earlier, Thom (1960) has analyzed 141 open-country stations averaging about 15
years of data per station and has published results for Fe(u) = 0.5, 0.98, and 0.99 which correspond to
mean times between occurrences of 2, 50, and 100 years, respectively.
Since he used the Type II distribution, the absolute values are not as accurate as we get from the
Type I. However, the trends are the same with either distribution and illustrate some very interesting
facts about the geographical variation of extreme winds.
The curve for the 2 year mean recurrence interval gives results which might be anticipated from
information on average winds presented earlier in the chapter. The High Plains have the highest
extreme wind speeds and the south-eastern and south-western United States have the lowest. This
relationship is basically the same as that for the average wind speeds. The 50 year and 100 year curves
are distinctly different, however. The once a century extreme wind in Western Kansas, Oklahoma, and
Texas is about 40 m/s, a figure exceeded in large areas of the coastal regions. The Gulf Coast and
Atlantic Coast, as far north as Southern Maine, have experienced extreme winds from tropical
cyclones, or hurricanes. The effect of these storms often extends inland from 100 to 200 miles. The
remainder of the United States experiences extreme winds largely from thunderstorms. This type of
storm accounts for over one third of the extreme-wind situations in the contiguous United States.
Water areas have a marked effect on extreme wind speeds. Where a location has unobstructed
access to a large body of water, extreme winds may be 15 m/s or more greater than a short distance
inland. High winds in cyclones and near water tend to remain steady over longer periods than for
thunderstorms. They also tend to be much more widespread in a given situation and hence cause
widespread damage, although damage to individual structures may not be greater than from
thunderstorm winds. The speed of the greatest gust experienced by a wind turbine will be somewhat
greater than the speed of the fastest mile because of the averaging which occurs during the period of
measurement. This will vary with the period chosen for the gust measurement and on the speed of the
fastest mile. It appears that a good estimate of a three second gust speed is 1.25 times the speed of the
fastest mile [Liu, H. (1982)]. That is, if the fastest mile is measured at 40 m/s, the fastest gust will
probably be about 1.25(40) = 50 m/s. This is the speed which should be used in talking about
survivability of wind turbines.
Standard civil engineering practice calls for ordinary buildings to be designed for a 50 year
recurrence interval, and for structures whose collapse do not threaten human safety to be designed for a
25 year recurrence interval. The civil engineers then typically add safety factors to their designs which
cause the structures to actually withstand higher winds. A similar approach would seem appropriate for
wind turbines [Liu, H. (1982)]. A 50 year recurrence interval would imply that a design for a maximum
gust speed of 50 m/s would be adequate over most of the United States, with perhaps 65 m/s being
desirable in hurricane prone areas.
Wind turbine costs tend to increase rapidly as the design speed is increased, so rather careful cost

| Wind Characteristics Pavesi - 91 -


studies are required to insure that the turbine is not over designed. Wind turbines are different from
public buildings and bridges in that they would normally fail in high winds without people getting hurt.
It may be less expensive to replace an occasional wind turbine with a design speed of 50 m/s than to
build all wind turbines to withstand 65 m/s.

| Wind Characteristics Pavesi - 92 -


4.7 Reference
ANSI Standard A58.1-1982, Minimum Design Loads for Buildings and other Structures, American
National Standards Institute, Inc., 1982.
Burr, I. W.: Applied Statistical Methods, Academic Press, New York, 1974.
Climatography of the United States, Series 82: Decennial Census of the United States Climate,
“Summary of Hourly Observations, 1951-1960” (Table B).
Cole, F. W.: Introduction to Meteorology, Wiley, New York, 1970.
Corotis, R. B.: Stochastic Modelling of Site Wind Characteristics, ERDA Report RLO/2342-77/2,
September 1977.
Corotis, R. B., A. B. Sigl, and J. Klein: “Probability Models of Wind Velocity Magnitude and
Persistence,” Solar Energy, Vol. 20, No. 6, 1978, pp. 483-493.
Corotis, R. B.: Statistic Models for Wind Characteristics at Potential Wind Energy Conversion Sites,
DOE Report DOE/ET/20283-1, January 1979.
Donn, W. L.: Meteorology, McGraw-Hill, New York, 1965.
Elliott, D. L.: Synthesis of National Wind Energy Assessments, Pacific Northwest Laboratory,
Richland, Wash., Report NTIS BNWL-2220 WIND-5, 1977.
Golding, E.: The Generation of Electricity by Wind Power, Halsted Press, New York, 1976.
Gumbel, E. J., Statistics of Extremes, Columbia University Press, New York, 1958.
Henry, David H. and Gary L. Johnson: “Distributions of Daily Extreme Winds and Wind Turbine
Operation,” IEEE Transactions on Power Apparatus and Systems, Vol. EC-1, No. 2, June, 1986,
pp. 125-130.
Hewson EW Wade JW (1977) Biological wind prospecting. Third Wind Energy workshop (CONF
770921), Washington DC, pp 335-348
Justus, C. G.: Winds and Wind System Performance, Franklin Institute Press, Philadelphia, 1978.
Justus, C. G., K. Mani, and A. Mikhail: Interannual and Month-to-Month Variations of Wind Speed,
DOE Report RLO-2439-78/2, April 1978.
Liu, H.: “Matching WECS Survivability with Severe Wind Characteristics: A Rational Approach,”
Wind and Solar Energy Conference, Kansas City, Mo., April 5-7, 1982.
Marrs RW Marwitz J (1977) Locating areas of high wind energy potential by remote observation of
eolian geomorphology and topography. Third Wind Energy workshop (CONF 770921),
Washington DC, pp307- 320
Odette, D. R.: A Survey of the Wind Energy Potential of Kansas, M.S. thesis, Electrical Engineering
Department, Kansas State University, Manhattan, Kans., 1976.
Panofsky, H. A.: “Wind Structure in Strong Winds below 150 m,” Wind Engineering, Vol. 1, No. 2,
1977, pp. 91-103.
Pasquill, F.: Atmospheric Diffusion, 2nd ed., Halsted Press, New York, 1974.
Putnam, P.C.: Power from the Wind, Van Nostrand, New York, 1948.
Riehl, H.: Introduction to the Atmosphere, McGraw- Hill, New York, 1965.
Simiu, E., M. J. Changery, and J. J. Filliben, Extreme Wind Speeds at 129 Stations in the Contiguous
United States, NBS Building Science Series 118, National Bureau of Standards, Washington, D.
C., March, 1979.
Spera, D. A., and T. R. Richards: “Modified Power Law Equations for Vertical Wind Profiles,”
Conference and Workshop on Wind Energy Characteristics and Wind Energy Sitting 1979,
Portland, Oreg., June 1979, Pacific Northwest Laboratory, Battelle Memorial Institute, Report

| Wind Characteristics Pavesi - 93 -


PNL-3214.
Thom, H. C. S.: “Distributions of Extreme Winds in the United States,” Journal of the Structural
Division, Proceedings of the ASCE, Vol. 86, No. ST4, April 1960, pp. 11-24.
Thom, H. C. S.: “New Distributions of Extreme Winds in the United States,” Journal of the Structural
Division, Proceedings of the ASCE, Vol. 94, No. ST7, Proc. Paper 6038, July 1968, pp. 1787-
1801.
Van der Hoven, I., (1957). ‘Power spectrum of horizontal wind speed in the frequency range from
0.0007 to 900 cycles per hour.’ J. Met., 14, 160–4.
WMO: World Meteorological Organization Guide to Meteorological Instrumentation and Observing
Practices, WMO-No. 8TP3, 4th ed., WHO, Geneva, 1971.

| Wind Characteristics Pavesi - 94 -


5 Equation Chapter (Next) Section 1 Aerodynamics of Wind Turbines
Wind turbine power production depends on the interaction between the rotor and the wind.
Experience has shown that the major aspects of wind turbine performance (mean power output and
mean loads) are determined by the aerodynamic forces generated by the mean wind. Periodic
aerodynamic forces caused by wind shear, off-axis winds and rotor rotation and randomly fluctuating
forces induced by turbulence and dynamic effects are the source of fatigue loads and are a factor in the
peak loads experienced by a wind turbine. These are, of course, important, but can only be understood
once the aerodynamics of steady state
operation have been understood.
A wind turbine extracts kinetic energy
from the wind by removing some of its
kinetic energy the wind must slow down
but only that mass of air which passes
through the rotor disc is affected.
Assuming that the affected mass of
air remains separate from the air which
does not pass through the rotor disc and
does not slow down a boundary surface can
be drawn containing the affected air mass
and this boundary can be extended
Figure 4.7.1 The Energy Extracting Stream-tube of a upstream as well as downstream forming a
Wind Turbine long stream-tube of circular cross section.
No air flows across the boundary and so the
mass flow rate of the air flowing along the stream-tube will be the same for all stream-wise positions
along the stream-tube. Because the air within the stream-tube slows down, but does not become
compressed, the cross-sectional area of the stream-tube must expand to accommodate the slower
moving air (Figure 4.7.1 The Energy Extracting Stream-tube of a Wind TurbineFigure 4.7.1).
Although kinetic energy is extracted from the airflow, a sudden step change in velocity is neither
possible nor desirable because of the enormous accelerations and forces this would require. Pressure
energy can be extracted in a step-like manner, however, and all wind turbines, whatever their design,
operate in this way. The presence of the turbine causes the approaching air, upstream, gradually to slow
down such that when the air arrives at the rotor disc its velocity is already lower than the free-stream
wind speed. The stream-tube expands as a result of the slowing down and, because no work has yet
been done on, or by, the air its static pressure rises to absorb the decrease in kinetic energy.
As the air passes through the rotor disc, by design, there is a drop in static pressure such that, on
leaving, the air is below the atmospheric pressure level. The air then proceeds downstream with
reduced speed and static pressure this region of the flow is called the wake. Eventually, far
downstream, the static pressure in the wake must return to the atmospheric level for equilibrium to be
achieved. The rise in static pressure is at the expense of the kinetic energy and so causes a further
slowing down of the wind. Thus, between the far upstream and far wake conditions, no change in static
pressure exists but there is a reduction in kinetic energy.

| Aerodynamics of Wind Turbines Pavesi - 95 -


5.1 The Actuator Disc Concept
The mechanism described above
accounts for the extraction of kinetic
energy but in no way explains what
happens to that energy; it may well be
put to useful work but some may be
spilled back into the wind as
turbulence and eventually be
dissipated as heat. Nevertheless, we
can begin an analysis of the
aerodynamic behaviour of wind
turbines without any specific turbine
design just by considering the energy
extraction process. The general device
that carries out this task is called an
actuator disc (Figure 5.1.1). The
analysis uses the following
assumptions:
 homogenous, incompressible,
steady state fluid flow
 no frictional drag
 an infinite number of blades
 uniform thrust over the disk or
rotor area
Figure 5.1.1 The Energy Extracting Stream-tube of a Wind
 a no rotating wake
Turbine
 the static pressure far upstream
and far downstream of the rotor is equal to the undisturbed ambient static pressure
Upstream of the disc the stream-tube has a cross-sectional area smaller than that of the disc and
an area larger than the disc downstream. The expansion of the stream-tube is because the mass flow
rate must be the same everywhere. The mass of air which passes through a given cross section of the
stream-tube in a unit length of time is AU, where  is the air density, A is the cross-sectional area and
U is the flow velocity. The mass flow rate must be the same everywhere along the stream tube and so
m  A  U   A d U d  A w U w (6.1.1)
The symbol  refers to conditions far upstream, d refers to conditions at the disc and w refers to
conditions in the far wake.
It is usual to consider that the actuator disc induces a velocity variation which must be
superimposed on the free-stream velocity. The stream-wise component of this induced flow at the disc
is given by -aU, where a is called the axial flow induction factor, or the inflow factor. At the disc,
therefore, the net stream-wise velocity is
Ud  U 1  a  (6.1.2)

The air that passes through the disc undergoes an overall change in velocity, U - Uw and a rate of

| Aerodynamics of Wind Turbines Pavesi - 96 -


change of momentum equal to the overall change of velocity times the mass flow rate:
Rate of Change of Momentum   U  U w  Ad Ud (6.1.3)
The force causing this change of momentum comes entirely from the pressure difference across
the actuator disc because the stream-tube is otherwise completely surrounded by air at atmospheric
pressure, which gives zero net force. Therefore,

p 
d  pd  A d   U   U w  A d U d
(6.1.4)
  U   U w  A d U  1  a 

To obtain the pressure difference p 


d  pd  Bernoulli’s equation is applied separately to the
upstream and downstream sections of the stream-tube; separate equations are necessary because the
total energy is different upstream and downstream. Bernoulli’s equation states that, under steady
conditions, the total energy in the flow, comprising kinetic energy, static pressure energy and
gravitational potential energy, remains constant provided no work is done on or by the fluid. Thus, for a
unit volume of air
1 2
U  p  gh  constant (6.1.5)
2
Upstream, therefore, we have
1 1
 U2  p   gh   Ud2  pd  d gh d (6.1.6)
2 2
Assuming the flow to be incompressible (=d) and horizontal (h=hd) then,
1 2 1
U  p  Ud2  pd (6.1.7)
2 2
Similarly, downstream,
1 2 1
U w  pw  Ud2  pd (6.1.8)
2 2
Subtracting these equations we obtain

p 
d  pd     U2  U 2w 
1
2
(6.1.9)

Equation (6.1.4) then gives

  U2  U2w  Ad   U  U w  Ad U 1  a 


1
(6.1.10)
2
and so
U w  1  2a  U (6.1.11)
That is, half the axial speed reduction in the stream-tube takes place upstream of the actuator disc

| Aerodynamics of Wind Turbines Pavesi - 97 -


and half downstream.
Equating the thrust values from Equations (6.1.4), (6.1.9), (6.1.11) and recognizing that the mass
flow rate is AdUd, one obtains:
U  U w
Ud 
2 (6.1.12)
 U  1  a 
Thus, the wind velocity at the rotor plane, using this simple model, is the average of the upstream
and downstream wind speeds. The quantity, Ua , is often referred to as the induced velocity at the
rotor, in which case velocity of the wind at the rotor is a combination of the free stream velocity and
the induced wind velocity. As the axial induction factor increases from 0, the wind speed behind the
rotor slows more and more. If a = 1/2, the wind has slowed to zero velocity behind the rotor and the
simple theory is no longer applicable.
The force on the air becomes, from equations (6.1.4) and (6.1.11)
F   pd  pd  A d
(6.1.13)
 2A d U 2 a 1  a 
An alternative control volume to the one in Figure 5.1.1 is shown in Figure 5.1.2.
On this alternative control volume there is a mass flow through the lateral boundary, since this is
not aligned with the streamlines. The axial momentum equation therefore becomes:

U 2w A w  U2  A  A w   mside U   U 2 A   F (6.1.14)

mside can be found from the conservation of mass:

U w A w  U  A  A w   mside  U A (6.1.15)


yielding:
mside  A w  U  U w  (6.1.16)

Figure 5.1.2 Alternative control volume around a wind turbine

| Aerodynamics of Wind Turbines Pavesi - 98 -


The conservation of mass also gives a relationship between Ad and Aw as:
m  Ud A d  U w A w (6.1.17)
Combining equations (6.1.16), (6.1.17) and (6.1.14) yields:
F  m  U  Uw 
 U d Ad  U   U w  (6.1.18)
 2Ad U (1  a) a
2

As this force is concentrated at the actuator disc the rate of work done by the force is FUd and
hence the power extraction from the air is given by
P  FUd  2Ad U3 a 1  a 
2
(6.1.19)
A power coefficient is then defined as
P
CP  (6.1.20)
1 3
U A
2  d
where the denominator represents the power available in the air, in the absence of the actuator disc.
Therefore,
CP  4a 1  a 
2
(6.1.21)
The force on the actuator disc caused by the pressure drop, given by Equation (6.1.13), can also
be non- dimensionalized to give a Coefficient of Thrust CT
F
CT 
1 2
U A (6.1.22)
2  d
 4a 1  a 
Note that the overall turbine efficiency is a function of both the rotor power coefficient and the
mechanical and electrical efficiency of the wind turbine:
P
  m el CP (6.1.23)
1
A d U3d
2

| Aerodynamics of Wind Turbines Pavesi - 99 -


5.2 The Betz Limit
The maximum value of CP occurs when
dCP
 4 1  a   8a 1  a 
2

da (6.2.1)
 4 1  a 1  3a   0
which gives a value of a =1/3.
Hence,
16
CP Max  0.593 (6.2.2)
27
The maximum achievable value of the power coefficient is known as the Betz limit after Albert
Betz the German aerodynamicist (1919). The limit is caused not by any deficiency in design, for, as
yet, we have no design, but because the stream-tube has to expand upstream of the actuator disc and so
the cross section of the tube where the air is at the full, free-stream velocity is smaller than the area of
the disc.
A problem arises for values of a≥½ because the wake velocity, given by (1- 2a)U, becomes zero,
or even negative; in these conditions the momentum theory, as described, no longer applies and an
empirical modification has to be made .
The variation of power coefficient and thrust coefficient with a is shown in Figure 5.2.1
In practice three effects lead to a decrease in the maximum achievable power coefficient:
 Rotation of the wake behind the rotor
 Finite number of blades and associated tip losses
 Non-zero aerodynamic drag

1.0
Uw/Uoo CT
.

0.8
U w /U oo

0.6
CP

0.4
C T

0.2
CP

0.0
0.0 0.2 0.4 0.6 0.8 1.0
a
Figure 5.2.1 Variation of CP, CT and UW/U with Axial Induction Factor a

| Aerodynamics of Wind Turbines Pavesi - 100 -


Experiments have
shown that the
assumptions of an ideal
wind turbine leading to
equation (6.1.22) are
only valid for an axial
induction factor, a, of
less than approximately
0.4. This is seen in
Figure 5.2.2, which
shows measurements of
CT as a function of a for
different rotor states. If
the momentum theory
were valid for higher
values of a, the velocity
in the wake would
become negative as can
readily be seen by
equation (6.1.11). As
Figure 5.2.2 The measured thrust coefficient CT as a function of the axial CT increases the
induction factor a and the corresponding rotor states expansion of the wake
increases and thus also
the velocity jump from U to Uw in the wake, see Figure 5.2.2. The ratio between the areas A and Aw
in Figure 5.1.1 can be found directly from the continuity equation as:

A
 1  2a (6.2.3)
Aw
For a wind turbine, a high thrust coefficient CT, and thus a high axial induction factor a, is present
at low wind speeds. The reason that the simple momentum theory is not valid for values of a greater
than approximately 0.4 is that the free shear layer at the edge of the wake becomes unstable when the
velocity jump U – Uw becomes too high and eddies are formed which transport momentum from the
outer flow into the wake.
This situation is called
the turbulent-wake state,
see Figure 5.2.2 and
Figure 5.2.3.
In the actuator disk
model, the power
extracted by the axial
force is (1-a)U·F.
However, if the same
Figure 5.2.3 Schematic view of the turbulent-wake state induced by the actuator disk, exerting a
unstable shear flow at the edge of the wake

| Aerodynamics of Wind Turbines Pavesi - 101 -


force F, is fixed on an airplane moving with speed U, the power required to move the disk would be
U·F. So it takes more power to drive the disk than the maximum power that can be generated by the
disk. This difference is understood when the flow around the actuator is also included in the analysis. It
then follows that the energy conversion by an actuator disk has an inherent dissipation of kinetic energy
into heat.
In the process of mixing between the far wake and the very far wake, the kinetic power in the
flow will not be conserved, but it will be partially converted into heat. This heat is generated by the
viscous force that accelerates the flow in the stream tube to the velocity V in the very far wake. In this
process the flow inside the stream tube gains less kinetic energy than the outer flow loses.
Let m be the indefinite but large mass flow in the wind, in which an actuator disk is placed
perpendicular to the flow direction. Only a fraction ε of m flows through the stream tube that just
encloses the actuator disk. In the far wake the kinetic power inside the stream tube is ½ε m (1-2a)2 U 2
and in the outer flow it is ½(1-ε) m U 2 . In the very far wake the kinetic power is ½ m V2. The
difference has to be the heat generated;
1
PHeat = 2 ṁ{[ε(1 − 2a)2 U∞
2 2]
+ (1 − ε)U∞ − V 2 } = −(1 − ε)aU∞ F (6.2.4)
Of course, this is also equal to ΔP-ΔPs. Where
1
Ps  m  U2  1  2a  U2    1  a  U F
2
(6.2.5)
2  

P  m  U2  V 2    1  a  U F
1
(6.2.6)
2
If we want to normalise to P= ½ρAd U 3 , the mass flow through the actuator disk (1-a)ρAU has
to be replaced by ε m . So we use P = ½ε m U 2 /(1-a) = FU/[4a(1-a)]. Since the mass flow through
the actuator disk is much smaller than the flow outside the wake, we take the limit ε→0, and find the
following power coefficients:
P
CT   4a 1  a  (6.2.7)
P
CP  4a 1  a 
2
(6.2.8)
PHeat
CHeat = = 4a2 (1 − a) (6.2.9)
P

Here CT refers to the transferred kinetic power, CP to the kinetic power actually extracted and
CHeat to the power in the viscous heating.
It follows that the maximum efficiency for the process of transfer of kinetic energy into useful
power by an actuator disk η is:
CP
  1 a (6.2.10)
CT
which is in agreement with Betz’s result. Our calculation makes clear that an actuator disk does not

| Aerodynamics of Wind Turbines Pavesi - 102 -


convert all transferred kinetic energy into useful energy. The energy balance reads:
CT=CP+CHeat (6.2.11)
As mentioned before, the maximum extractable useful power from the flow is obtained for a=1/3.
In that case a fraction CT = 24/27 of the flow of kinetic energy P is transferred. From this, 2/3 is
extracted as useful power, and 1/3 is dissipated as heat. Figure 5.2.4 shows schematically the power
transfer by an actuator disk representing a wind turbine. We introduced Ui = -aU for the induction
velocity in order to make the model more general, so that the situation for an actuator disk representing
a propeller is also included.
For an actuator disk
representing a rotor in hover
(U=0) it follows from
equations (6.2.7) to (6.2.9)
and from Figure 5.2.4 that the
power required to yield F is
Ui·F and that this power is
entirely converted into heat. It
first turns up as kinetic
Figure 5.2.4 Schematic view of the kinetic energy transfer by an energy, which is eventually
actuator disk. dissipated via turbulent
mixing and viscous shear as
heat. For an actuator in propeller state (U≥0 and Ui≥0), the engine power is (U+Ui)·F and the heat
produced is Ui·F and the kinetic energy of the flow is increased by U·F. (The co-ordinate system is still
attached to the actuator).
We conclude that the inherent limitation to the efficiency of energy extraction by an actuator disk
is determined by dissipation as heat. This dissipation is a/(1-a) times the extracted useful energy. The
heat capacity of the mass flow through a wind turbine is so large that the heat generated will hardly
affect the temperature. To give an example: a wind turbine operating at 10 m/s at the Lanchester-Betz
limit will transfer 44.4J of kinetic energy per unit of air mass into 29.6J of useful work and 14.8J of
heat. This heat raises the temperature by only 0.015°C.
In practice it will be even less since the heat generated is not limited to the flow inside the stream
tube.

| Aerodynamics of Wind Turbines Pavesi - 103 -


5.3 Equation Section (Next) Rotor Disc
Theory
The manner in which the extracted energy is
converted into usable energy depends upon the
particular turbine design. Most wind energy
converters employ a rotor with a number of blades
rotating with an angular velocity  about an axis
normal to the rotor plane and parallel to the wind
direction.
The blades sweep out a disc and by virtue of
their aerodynamic design develop a pressure
difference across the disc, which, as discussed in
the previous section, is responsible for the
diminution of axial momentum in the wake.
Associated with the axial momentum decrease is a
energy reduction which can be collected by, say,
an electrical generator attached to the rotor shaft
if, as well as a thrust, the rotor experiences a
torque in the direction of rotation.
The generator exerts a torque equal and
opposite to that of the airflow which keeps the
rotational speed constant. The work done by the
aerodynamic torque on the generator is converted
into electrical energy. Figure 5.3.1 The Trajectory of an Air Particle
The exertion of a torque on the rotor disc by Passing Through the Rotor Disc
the air passing through it requires an equal and
opposite torque to be imposed upon the air. The
consequence of the reaction torque is to cause the
air to rotate in a direction opposite to that of the
rotor; the air gains angular momentum and so in
the wake of the rotor disc the air particles have a
velocity component in a direction which is
tangential to the rotation as well as an axial
component (Figure 5.3.1).
The acquisition of the tangential component
of velocity by the air means an increase in its
kinetic energy which is compensated for by a fall
in the static pressure of the air in the wake in
addition to that which is described in the previous
section.
The flow entering the actuator disc has no
rotational motion at all. The flow exiting the disc
does have rotation and that rotation remains Figure 5.3.2 Tangential Velocity Grows Across
constant as the fluid progresses down the wake. the Disc Thickness

| Aerodynamics of Wind Turbines Pavesi - 104 -


The transfer of rotational motion to the air takes place entirely across the thickness of the disc (see
Figure 5.3.2). The change in tangential velocity is expressed in terms of a tangential flow induction
factor a’. Upstream of the disc the tangential velocity is zero. Immediately downstream of the disc the
tangential velocity is 2ra’. At the middle of the disc thickness, a radial distance r from the axis of
rotation, the induced tangential velocity is ra’. Because it is produced in reaction to the torque the
tangential velocity is opposed to the motion of the rotor.
An abrupt acquisition of tangential velocity cannot occur in practice. Figure 5.3.2 shows the flow
accelerating in the tangential direction as it is ‘squeezed’ between the blades; the separation of the
blades has been reduced for effect but it is the increasing solid blockage that the blades present to the
flow as the root is approached that causes the high values of tangential velocity close to the root.
The tangential velocity will not be the same for all radial positions and it may well also be that the
axial induced velocity is not the same. To allow for variation of both induced velocity components
consider only an annular ring of the rotor disc which is of radius r and of radial width r.
The increment of rotor torque acting on the annular ring will be responsible for imparting the
tangential velocity component to the air whereas the axial force acting on the ring will be responsible
for the reduction in axial velocity. The whole disc comprises a multiplicity of annular rings and each
ring is assumed to act independently in imparting momentum only to the air which actually passes
through the ring.
The torque on the ring will be equal to the rate of change of angular momentum of the air passing
through the ring. Thus,
M  Ad U  1  a  2a 'r 2 (6.3.1)

where Ad is taken as being the area of an annular ring.


The driving torque on the rotor shaft is also M and so the increment of rotor shaft power output
is
P  M (6.3.2)
The total power extracted from the wind by slowing it down is therefore determined by the rate of
change of axial momentum given by Equation (6.1.19).

P  2Ad U3 a 1  a 
2
(6.3.3)
Hence
2Ad U3 a 1  a   2A d U  1  a  2a 'r 2
2
(6.3.4)
and
U 2 a 1  a   2 r 2a ' (6.3.5)

r is the tangential velocity of the spinning annular ring and so  r  r U  is called the local speed
ratio. At the edge of the disc r = R and   R U  is known at the tip speed ratio. Thus

a 1  a    2ra ' (6.3.6)

The area of the ring is A d  2rr therefore the incremental shaft power is, from Equation

| Aerodynamics of Wind Turbines Pavesi - 105 -


(6.3.1),
1 
P  dM   U3 2rr  4a ' 1  a   2r (6.3.7)
2 
The term in brackets represents the power flux through the annulus, the term outside the brackets,
therefore, is the efficiency of the blade element in capturing the power, or blade element efficiency:
r  4a ' 1  a   2r (6.3.8)
In terms of power coefficient
d 4U3 1  a  a '  2r r 8 1  a  a '  2r r
CP   (6.3.9)
dr 1 3 2
U  R R2
2
d
CP  8 1  a  a '  23 (6.3.10)
d
where  =r/R.
Knowing how a and a’ vary radially, equation (6.3.10) can be integrated to determine the overall
power coefficient for the disc for a given tip speed ratio, .

Maximum power
The values of a and a’ which will provide the maximum possible efficiency can be determined by
differentiating Equation (6.3.8) by either factor and putting the result equal to zero. Whence
d 1 a
a (6.3.11)
da ' a'
From Equation (6.3.6)
d 2
a r (6.3.12)
da ' 1  2a
giving
a '  2r  1  a 1  2a  (6.3.13)
The combination of Equations (6.3.6) and (6.3.11) gives the required values of a and a’ which
maximize the incremental power coefficient:
1 a(1−a)
a and a′ = (6.3.14)
3 λ2r

The axial flow induction for maximum power extraction is the same as for the non rotating wake
case, that is, a=1/3 and is uniform over the entire disc. On the other hand a’ varies with radial position.
From Equation (6.3.10) the maximum power is

| Aerodynamics of Wind Turbines Pavesi - 106 -


1
CP   8 1  a  a '  23d (6.3.15)
0

Substituting for the expressions in equations (6.3.14)


1
 a 1  a  
CP   8 1  a   2 2   23d
0    (6.3.16)
16
 4a 1  a  
2

27
Which is precisely the same as for the non-rotating wake case.

| Aerodynamics of Wind Turbines Pavesi - 107 -


5.4 Wake structure
The angular momentum imparted to the wake increases the kinetic energy in the wake but this
energy is balanced by a static pressure decrease:
1
p  U2  2a 'r 
2
(6.4.1)
2
Substituting the expression for a’ given by equations (6.3.14)

 a 1  a  
2
1
p  U 2  2  (6.4.2)
2   
The tangential velocity increases with decreasing radius (equation (6.3.14)) and so the pressure
decreases creating a radial pressure gradient. The radial pressure gradient balances the centrifugal force
on the rotating fluid. The pressure drop across the disc caused by the rate of change of axial momentum
as developed (equation (6.1.13)) is additional to the pressure drop associated with the rotation of the
wake and is uniform over the whole disc.
If the wake did not expand as it slows down the rotational wake structure together with the
rotational pressure gradient would not change as the wake develops whereas the pressure loss caused
by the change of axial momentum will gradually reduce to zero in the fully-developed wake, as shown
in Figure 5.2.3. The pressure in the fully developed wake would therefore be atmospheric
superimposed on which would be the pressure loss given by equation (6.4.2). Consequently, the axial
force on the fluid in the wake causing it to slow down would be only that caused by the uniform
pressure drop across the disc given by equation (6.1.13), as is assumed in the simple theory of Section
0. The rotational pressure drop does not contribute to the change of axial momentum.
In fact, the wake expands and the full details of the analysis are given by Glauert (1935a).
Glauert’s analysis is applied to propellers where the flow is accelerated by the rotor but this is only a
matter of reversing the signs of the flow induction factors. The inclusion of flow expansion and wake
rotation in a fully integrated momentum theory shows that the axial induced velocity in the developed
wake is greater than 2a but the effect is only significant at tip speed ratios less than about 1.5, which is
probably outside of the operating range for most modern wind turbines. The analysis does, however,
demonstrate that the kinetic energy of wake rotation is accounted for by reduced static pressure in the
wake. Glauert’s conclusion about wake expansion and its interaction with wake rotation is that its
inclusion makes little difference to the results obtained from the simple axial momentum theory and so
can be ignored. Where, in the same reference, Glauert deals with ‘Windmills and Fans’ (1935b) he
adopts the simple momentum theory but then has to account for kinetic energy of wake rotation, which
he does by assuming that it is drawn from the kinetic energy of the flow. The rotational kinetic energy
of the wake is therefore regarded as a loss and reduces the level of the energy that can be extracted.
Consequently, at low local speed ratios, the inboard sections of a rotor, the local aerodynamic
efficiency falls below the Betz limit. Most authors since Glauert have assumed the same conclusion
but, in fact, Glauert himself has demonstrated that the conclusion is wrong. The error makes very little
difference to the final results for most modern wind turbines designed for the generation of electricity.
For wind pumps, where a high starting torque and high solidity are required, the error would probably
be very significant because they operate at very low tip speed ratios.

| Aerodynamics of Wind Turbines Pavesi - 108 -


5.5 Shrouted Rotors
It is possible to exceed the Betz limit by
placing the wind turbine in a diffuser. If the
cross-section of the diffuser is shaped like an
aerofoil, a lift force will be generated by the
flow through the diffuser as seen in Figure
5.2.1.
As shown in de Vries (1979), the effect of
this lift is to create a ring vortex, which by the
Biot-Savart law will induce a velocity to
increase the mass flow through the rotor. The
axial velocity in the rotor plane is denoted by
Ud,d, and ad is the induced factor related to the
ratio between Ud,d and the wind speed U, i.e. Figure 5.5.1 Ideal flow through a wind turbine in a
1-ad = Ud,d/U. A 1-D analysis of a rotor in a diffuser
diffuser gives the following expression for the power coefficient:
P
CP,d 
1 3
U  A d
2
F
 (6.5.1)
1 2 U
U  Ad
2 U d,d
 CT 1  a d 
For an ideal bare turbine equations CP=4a(1-a)2 and CT=4a(1-a) are valid yielding:
CP =CT(1-a) (6.5.2)
Combining equations (6.5.1) and (6.5.2) yields:
CP,d 1  a d
 (6.5.3)
CP 1 a
Further, the following equations are valid for the mass flow through a bare turbine mb and the
mass flow through a turbine in a diffuser md:
m  1  a  U  A d

U A d U  A d (6.5.4)
 1 a
m U d,d A d

U  A d U  A d (6.5.5)
 1 ad

| Aerodynamics of Wind Turbines Pavesi - 109 -


Combining equations (6.5.3), (6.5.4) and (6.5.5) yields:
CP,d md
 (6.5.6)
CP m
Equation (6.5.6) states that the relative
increase in the power coefficient for a
shrouded turbine is proportional to the ratio
between the mass flow through the turbine in
the diffuser and the same turbine without the
diffuser. Equation (6.5.6) is verified by
computational fluid dynamics (CFD) results as
seen in Figure 5.5.2, where for a given
geometry the computed mass flow ratio md/mb
is plotted against the computed ratio Cp,d /Cp,b.
The CFD analysis is done on a simple
geometry, without boundary layer bleed slots,
as suggested by Gilbert and Foreman (1983).
To check this approach some initial
computations were made without the diffuser
Figure 5.5.2 Computed mass flow ratio plotted and in Figure 5.5.2, which shows the
against the computed power coefficient ratio for a relationship between the thrust and power
bare and a shrouded wind turbine, respectively coefficients, it is seen that this approach gave
good results compared to the following
theoretical expression, which can be derived from equations CP=4a(1-a)2 and CT=4a(1-a):
1

CP,b  CT 1  1  CT
2
 (6.5.7)

In Figure 5.5.3 it is also seen that


computations with a wind turbine in a diffuser
gave higher values for the power coefficient than
the Betz limit for a bare turbine.
The results are dependent on the actual
diffuser geometry, in other words on the amount
of lift which can be generated by the diffuser. An
adverse pressure gradient is present for the flow in
the diffuser, and the boundary layer will separate
if the ratio between the exit area and the area in
the diffuser becomes too high. To increase the lift,
giving a higher mass flow through the turbine and
thus a higher power output, any trick to help
prevent the boundary layer from separating is
allowed, for example vortex generators or
Figure 5.5.3 Computed power coefficient for a boundary bleed slots. The computations in Figure
rotor in a diffuser as a function of the thrust 5.5.3 and the wind tunnel measurements of Gilbert
coefficient CT and Foreman (1983) show that the Betz limit can

| Aerodynamics of Wind Turbines Pavesi - 110 -


be exceeded if a device increasing the mass flow through the rotor is applied, but this still has to be
demonstrated on a full size machine. Furthermore, the increased energy output has to be compared to
the extra cost of building a diffuser and the supporting structure.

| Aerodynamics of Wind Turbines Pavesi - 111 -


5.6 Equation Section (Next)Vortex Cylinder Model of the Actuator Disc
In inviscid flow, the flow field is determined by the Euler equation which describes the
interaction between pressure distribution, external forces and velocity field and assumes that no internal
friction (viscosity) exists. When an aerodynamic object is placed in a fluid in motion a pressure
distribution over the surface of the object comes into being. This pressure distribution is in agreement
with the velocity field around the object. The words ‘in agreement with’ were used to emphasise the
mutual interaction between pressure distribution on the object and velocity field around the object
instead of a causal connection. In summary: an object in combination with a flow causes a combination
of a velocity field and a pressure distribution. The resulting velocity field can be described as the sum
of the undisturbed fluid motion and the motion described by a vorticity distribution.
Vortices describe induced velocities, but only a specific portion of them are ‘induction velocities’
in Prandtl-terms. This portion accounts for the difference between the three-dimensional steady or
unsteady practical situations and the 'twodimensional steady' wind tunnel situation. The difference
consists in general of three types of vortices,
shown in Figure 5.6.1. The first is the trailing
vorticity of the tips of a finite airfoil (what
BT induces at BB); the second the vorticity
shed from the airfoil when the bound
circulation changes over time (what PP
induces at BB); and the third is the absence
of the (shed) vorticity outside the span of the
airfoil (what SP, which is the shed vorticity
of AB, induces at BB. This contribution does
not exist in the 3D-situation, but is present in
the 2D-situation. So the difference has to be
corrected). For a precise definition of the
third contribution we refer to Van Holten. It
Figure 5.6.1 Prandtl’s induction velocities. is especially important for helicopter rotors
with their strong variations in circulation
with azimuth. In the case of a wind turbine it
is normally sufficient to account for the tip
vortices only. It should be noted that the
velocity pattern described by the bound
vorticity that was present in the wind tunnel
is excluded from the induction velocities in
'Prandtl terms'.
The linear momentum theory uses the
concept of the actuator disc across which a
pressure drop develops constituting the
energy extracted by the rotor. The analysis
can be extended to the case where the
rotating rotor generates angular momentum,
which can be related to rotor torque. In the
Figure 5.6.2 Flow visualisation with smoke, revealing
case of a rotating wind turbine rotor, the flow
the tip vortices

| Aerodynamics of Wind Turbines Pavesi - 112 -


behind the rotor rotates in the opposite direction to the rotor, in reaction to the torque exerted by the
flow on the rotor. An annular stream tube model of this flow, illustrating the rotation of the wake, is
shown in Figure 5.6.3. In the rotor disc theory the actuator disc is depicted as being swept out by a
multiplicity of aerofoil blades each with
radially uniform bound circulation . From
the tip of each blade a helical vortex of
strength  converts downstream with the
local flow velocity (Figure 5.6.4). If the
number of blades is assumed to be very large
but the solidity of the total is finite and small
then the accumulation of helical tip vortices
will form the surface of a tube. As the
number of blades approaches infinity the tube
surface will become a continuous tubular
vortex sheet.
From the root of each blade, assuming
Figure 5.6.3 Flow visualisation with smoke, revealing it reaches to the axis of rotation, a line vortex
smoke trails being ‘sucked’ into the vortex spirals. of strength  will extend downstream along
the axis of rotation contributing to the total
root vortex of strength . The vortex tube
will expand in radius as the flow of the wake
inside the tube slows down. Vorticity is
confined to the surface of the tube, the root
vortex and to the bound vortex sheet swept
by the multiplicity of blades to form the rotor
disc; elsewhere in the wake and everywhere
else in the entire flow field the flow is
irrotational.
The nature of the tube’s expansion
cannot be determined by means of the
momentum theory and so, as an
Figure 5.6.4 Helical Vortex Wake Shed by Rotor with approximation, the tube is allowed to remain
Three Blades Each with Uniform Circulation  cylindrical Figure 5.6.5. The Biot–Savart law
is used to determine the induced velocity at
any point in the vicinity of the actuator disc.
The cylindrical vortex model allows the
whole flow field to be determined and is
accurate within the limitations of the non-
expanding cylindrical wake.
The generation of rotational kinetic
energy in the wake results in less energy
extraction by the rotor than would be
expected without wake rotation. In general,
the extra kinetic energy in the wind turbine
Figure 5.6.5 Simplified Helical Vortex Wake Ignoring wake will be higher if the generated torque is
Wake Expansion

| Aerodynamics of Wind Turbines Pavesi - 113 -


higher.
The vortex cylinder has surface vorticity which follows a
helical path with a helix angle  or, as it has been termed
previously, the flow angle at the blade tip. The strength of the
vorticity is g  d dn , where n is a direction in the tube surface
normal to the direction of , and has a component g   g cos 
2 r
parallel to the rotor disc. Due to g the axial (parallel to the axis of
rotor rotation) induced velocity at the rotor plane is uniform over
the rotor disc and can be determined by means of the Biot–Savart
law as
g
Figure 5.6.6 The Geometry of ud    aU (6.6.1)
2
the Vorticity in the Cylinder
Surface In the far wake the axial induced velocity is also uniform
within the cylindrical wake and is
u w  g   2aU  (6.6.2)
The ratio of the two induced velocities corresponds to that of the simple momentum theory and
justifies the assumption of a cylindrical vortex sheet.
The total circulation on all of the multiplicity of blades is  which is shed at a uniform rate into
the wake in one revolution. So, from Figure 5.6.6 in which the cylinder has been slit longitudinally and
opened out flat,

g (6.6.3)
2r sin 
hence
 cos   r 1  a '
g   (6.6.4)
2r sin  2r U (1  a)
therefore
 cos   r 1  a '
2aU   (6.6.5)
2r sin  2r U (1  a)
So, the total circulation is related to the induced velocity
4aU2 a(1  a)
 (6.6.6)
 1  a '
Just as a vortex is shed from each blade tip a vortex is also shed from each blade root. If it is
assumed that the blades extend to the axis of rotation, obviously not a practical option, then the root
vortices will each be a line vortex running axially downstream from the centre of the disc. The
direction of rotation of the all the root vortices will be the same forming a core, or root, vortex, of total
strength . The root vortex is primarily responsible for inducing the tangential velocity in the wake

| Aerodynamics of Wind Turbines Pavesi - 114 -


flow and in particular the tangential velocity on the rotor disc.
On the rotor disc surface the tangential velocity induced by the root vortex, given by the Biot–
Savart law, is

ra ' 
4r (6.6.7)

a'
4r 2
This relationship can also be derived from the momentum theory: the rate of change of angular
momentum of the air which passes through an annulus of the disc of radius r and radial width dr is
equal to the torque increment imposed upon the annulus
dM  U 1  a  2 dr 2a ' r 2 (6.6.8)
By the Kutta–Joukowski theorem the lift per unit radial width is
L   w   (6.6.9)

where ( w   ) is a vector product.


d
M  w r sin 
dr (6.6.10)
 r U  1  a 
Equating the two expressions gives

a' (6.6.11)
4r 2
Hence
U 2 a 1  a 
a' 2 2
 r 1  a ' 
(6.6.12)
a 1  a 
 2
 r 1  a '
so
a 1  a 
a ' 1  a '  (6.6.13)
 2r
Equation (6.6.13) is not quite the same as equation (6.3.14) and it can be shown that this is a
result of ignoring the wake expansion.
The torque on an annulus of radius r and radial width r is

| Aerodynamics of Wind Turbines Pavesi - 115 -


d
Mr  wr sin r
dr
4rU3 a 1  a 
2

 r (6.6.14)
 1  a '
dM 1 8rU  a 1  a 
3 2


dr 2  1  a '
Power
dP dM

dr dr
1 8rU  a 1  a 
3 2

 (6.6.15)
2 1  a '
1 4R U  a 1  a 
2 3 2

P
2 1  a '
Power coefficient
4a 1  a 
2

CP  (6.6.16)
1  a '
The reduced efficiency compared with the simple actuator disc result, CP=4a(1-a)2, is caused by
the energy required to spin the wake, as a rigid body, with an angular velocity 2a 't  . It should be noted
that any additional rotational energy is accounted for by the loss of static pressure caused by the
pressure gradient which balances the centrifugal forces on the rotating fluid.
The general momentum theory of Glauert (1935a) includes the expansion of the wake and shows
that no contribution at all is actually required from the kinetic energy of the free-stream flow to
maintain wake rotation; all the kinetic energy of rotation is derived from static pressure energy.
We can also analyse the vortex cylindrical model by using a control volume that moves with the
angular velocity of the blades, the energy equation can be applied in the sections before and after the
blades to derive an expression for the pressure difference across the blades (Glauert, 1935). Note that
across the flow disk, the angular velocity of the air relative to the blade increases from r to r+2a’r
(Figure 5.3.2), while the axial component of the velocity remains constant. The rothalpy is constant so:

p 
d  pd   r 2 2     2a ' 
1
2 
2
 (6.6.17)
 2a ' 1  a ' 2 r 2
The resulting thrust on an annular element, dF , is:
dF   pd  pd  2rdr
(6.6.18)
 4a ' 1  a '  2 r 3dr

| Aerodynamics of Wind Turbines Pavesi - 116 -


Following the previous linear momentum analysis, the thrust on an annular cross-section can also
be determined by the following expression that uses the axial induction factor, a,
dF  4a 1  a  U2 rdr (6.6.19)
Equating the two expressions for thrust gives:
a 1  a 
 2r 
a ' 1  a '
That is the same result of equation (6.6.13)
Next, one can derive an expression for the torque on the rotor by applying the conservation of
angular momentum. For this situation, the torque exerted on the rotor, M, must equal the change in
angular momentum of the wake. On an incremental annular area element this gives:
dM  dm    2a '   r
 U  1  a  2rdr   2a '  r

Using the definition of the local tip speed ratio,  r  r U  and the tip speed ratio   R U 
the expression for the torque generated at each element becomes:
U3
dM  4R 2 a ' 1  a  3rd r (6.6.20)
 2
The power generated at each element, dP , is given by:

U3
dP  4R 2 a ' 1  a  3rd r (6.6.21)
2
It can be seen that the power from any annular ring is a function of the axial and angular
induction factors and the tip speed ratio. The axial and angular induction factors determine the
magnitude and direction of the airflow at the rotor plane. The local speed ratio is a function of the tip
speed ratio and radius. The incremental contribution to the power coefficient, CP,

8
a ' 1  a  3r d r
 2 0
CP  (6.6.22)

In order to integrate this expression, one needs to relate the variables a , a', and r, (see Glauert
1948, Sengupta and Verma, 1992). Solving Equation (6.6.13) to express a' in terms of a, one gets:

1 1  4 
a'   1  2 a 1  a   (6.6.23)
2 2  r 
The aerodynamic conditions for the maximum possible power production occur when the term
a’(1-a) in equation (6.6.23) is at its greatest value. Substituting the value for a' from equation (6.6.23)
into a'(1- a) and setting the derivative with respect to a equal to zero yields:

| Aerodynamics of Wind Turbines Pavesi - 117 -


1  a  4a 1
2

 2
 (6.6.24)
1  3a
r

This equation defines the axial induction factor for maximum power as a function of the local tip
speed ratio in each annular ring (Figure 5.6.7). Substituting into equation (6.6.13), one finds that, for
maximum power in each annular ring:
1  3a
a' (6.6.25)
4a  1
If equation (6.6.24) is differentiated with respect to a , one obtains a relationship between dr, and

0.7
r=0.50
0.6
a'
0.5

0.4
r=0.75
0.3
r=1.00
0.2
r=1.50 r=2.00 r=3.00
r=3.00
0.1

0
0 0.2 0.4 0.6 0.8 1
a
Figure 5.6.7 Tangential induction factor versus Axial induction factor
.

6.0

5.0
a'opt  r opt

1/4 1/3
4.0

3.0
a'opt  r opt
2.0

1.0

0.0
0.23 0.25 0.27 0.29 0.31 0.33 0.35
a
Figure 5.6.8 Tangential optimum induction factor versus axial induction
factor
| Aerodynamics of Wind Turbines Pavesi - 118 -
da at those conditions that result in maximum power production:

2 r d r  6  4a  11  2a  / 1  3a   da
2 2
(6.6.26)
 
Now, substituting the Equations (6.6.24)-(6.6.26) into the expression for the power coefficient
(Equation (6.6.22)) gives:

24  1  a 1  2a 1  4a  
Tip a

CP  2   da (6.6.27)
 a Hub  1  3a  
Were the lower limit of integration, aHub , corresponds to axial induction factor for r = 0 and the
upper limit, aTip, corresponds to the axial induction factor at r = . Also, from equation (6.6.24):

1  a 1  4a 
2

 
2 Tip Tip
(6.6.28)
1  3a 
Tip

Note that from equation (6.6.24), aHub = 0.25 gives r=0. Equation (6.6.28) can be solved for the
values of aTip that correspond to operation at tip speed ratios of interest (Figure 5.6.8). Note also from
equation (6.6.28), aTip = 1/3 is the upper limit of the axial induction factor, a , giving an infinitely large
tip speed ratio r=. Therefore a’ decreases increasing r that is increasing the radius and it is zero
when the blade has an infinite length or the wind velocity is zero whereas it tends to  on the axis of
the turbine (r=r=0). These consideration mean that the vorticity of the flow field and the dissipation
connected with it becomes high decreasing the radius whereas it decrease at the tip of the blade. This
consideration induces to design turbines with the hub radius with relative high values removing the
internal part that little contribute to the overall power.
It is interesting to note that a=1/3 correspond to the irritation flow field of the Actuator Disc
Theory.
The definite integral can be evaluated by changing variables: substituting x for (I - 3a) in equation
(6.6.27). The result is (Eggleston and Stoddard. 1987):
0.25
8  64 5 1 
  x  72x  124x  38x  63x  12 ln  x    4x 
4 3 2
CP (6.6.29)
Max
729  5  13a Tip

The results of this analysis are graphically represented in Figure 5.6.9, which also shows the Betz
limit of the ideal turbine based on the previous linear momentum analysis. The results show that, the
higher the tip speed ratio, the greater the maximum theoretical CP.

| Aerodynamics of Wind Turbines Pavesi - 119 -


0.7

.
Betz Limit
0.6

CP
0.5

0.4

0.3

0.2

0.1

0.0
0 2 4 6 8 10

Figure 5.6.9 Theoretical maximum power coefficient as a
function of tip speed ratio for an ideal horizontal axis wind turbine,
with and without wake rotation

| Aerodynamics of Wind Turbines Pavesi - 120 -


5.7 Axial flow field
By means of the
Biot-Savart law the
induced velocity in the
wind-wise (axial)
direction at the
actuator disc can be
determined both
upstream of the disc
and downstream in the
developing wake, as
well as on the disc
itself. The flow field
(net velocity) is
axisymmetric and a
radial cross section is Figure 5.7.1 The Radial and Axial Variation of Axial Velocity in the
shown in Figure 5.7.1. Vicinity of an Actuator Disc, a=1/3
Both radial and axial
distances are divided by the disc radius with the axial distance being measured downstream from the
disc and the radial distance being measured from the rotational axis. The velocity is divided by the
wind speed.
The axial velocity within the wake is sharply lower than without and is radially uniform at the
disc and in the far wake, just as the momentum theory predicts.
There is a small acceleration of the flow
immediately outside of the wake. The induced
velocity on the wake cylinder itself is 1/2 a at
the disc and a in the far wake.

5.8 Tangential flow field


The tangential induced velocity is
determined not only by the root vortex but
also by the component of vorticity g sin on
the wake cylinder and the bound vorticity on
the rotor disc. At a radial distance equal to half
the disc radius, as an example, the axial
variation of the three contributions are shown
in Figure 5.8.1
The bound vorticity causes rotation in
opposite senses upstream and downstream of
the disc with a step change across the disc.
The upstream rotation, which is in the same
sense as the rotor rotation, is nullified by the Figure 5.8.1 The Axial Variation of Tangential
root vortex, which induces rotation in the Velocity in the Vicinity of an Actuator Disc at 50%
opposite sense to that of the rotor. The Radius, a =1/3, =6

| Aerodynamics of Wind Turbines Pavesi - 121 -


downstream rotation is in the same sense for
both the root vortex and the bound vorticity the
stream-wise variations of the two summing to
give a uniform velocity in the stream-wise
sense.
The vorticity located on the surface of the
wake cylinder makes a small contribution. At
the disc itself the bound vorticity induces no
rotation, see Figure 5.8.1, the wake cylinder
induces no rotation either and so it is only the
root vortex which does induce rotation and that
value is half the total induced generally in the
wake. It is now clear why only half the
rotational velocity is used to determine the flow
angle at the disc.
The rotational flow is confined to the
Figure 5.8.2 The Axial Variation of Tangential wake, that is, inside the cylinder. There is no
Velocity in the Vicinity of an Actuator Disc at 101 rotational flow anywhere outside of the wake,
percent Radius, a =1/3, =6 neither upstream of the disc or outside the
cylinder. The rotational velocity within the
cylinder falls with increasing radius but is not zero at the outer edge of the wake, therefore there is an
abrupt fall of rotational velocity across the cylindrical surface. The contributions of the three vorticity
sources to the rotational flow at a radius of 101 percent of the disc radius are shown in Figure 5.8.2; the
total rotational flow is zero at all axial positions but the individual components are not zero.

5.9 Radial flow field


Although the vortex cylinder theory has been simplified by not allowing the cylinder to expand,
nevertheless the theory predicts flow expansion. A radial velocity distribution is predicted by the theory
as shown in Figure 5.9.1 which shows a longitudinal section of the flow field through the rotor disc.
The radial velocity, as calculated, is greatest as the flow passes through the rotor disc and,
although not shown in Figure 5.9.1, it is infinite at the edge of the disc. The situation is very similar to
the determination of the
potential flow field around a
flat, solid, circular disc
which is normal to the
oncoming flow; an infinite
radial velocity is predicted
at the disc edge for the flow
to pass around and continue
radially inwards on the
downstream side. In
practice there would be
insufficient static pressure
available to fuel an infinite,
or even a very high, velocity
Figure 5.9.1 Flow Field Through an Actuator Disc for a=1/3

| Aerodynamics of Wind Turbines Pavesi - 122 -


and so some discontinuity in the flow must occur. The presence of even the smallest amount of
viscosity would produce a thick boundary layer towards the disc edge because of the high radial
velocity. The viscosity in the boundary layer would absorb much of the available energy and dissipate
it as heat so that as the flow accelerated around the disc edge the maximum velocity attainable would
be limited by the static pressure approaching zero. Instead of the flow moving around the edge it would
separate from the edge and continue downstream leaving a very low pressure region behind the disc
with very low velocity—a stagnant region.
In the case of the permeable rotor disc there would be some flow through the disc, which would
behave as described above, but the separation of the flow at the disc edge would produce an additional
low pressure in the wake. The problem of the infinite radial velocity at the rotor disc edge arises
because of the assumption of an infinite number of rotor blades. If the theory is modified such that
there are only a few blades the infinite radial velocity disappears. However, if, for a given rotor, the tip
speed ratio is increased, with a consequent increase of the axial flow induction factor, the radial
velocity at the tip rises sharply and the problem of edge separation returns, which is what actually
occurs.

5.10 Conclusions
Despite the exclusion of wake expansion, the vortex theory produces results largely in agreement
with the momentum theory and enlightens understanding of the flow through an energy extracting
actuator disc.

| Aerodynamics of Wind Turbines Pavesi - 123 -


6 Rotor Blade Theory
6.1 2-D Blade Aerodynamics
Wind turbine blades are
long and slender structures
where the spanwise velocity
component is much lower than
the streamwise component, and
it is therefore assumed in many
aerodynamic models that the Figure 6.1.1 Schematic view of streamlines past an airfoil
flow at a given radial position is
two dimensional and that 2-D aerofoil data can thus be
applied. Two-dimensional flow is comprised of a plane,
as shown in Figure 6.1.1.
In order to realize a 2-D flow it is necessary to
extrude an aerofoil into a wing of infinite span. On a real
wing the chord and twist changes along the span and the
wing starts at a hub and ends in a tip, but for long slender
wings, like those on modern gliders and wind turbines.
Figure 6.1.1 shows the leading edge stagnation point
present in the 2-D flow past an aerofoil. The reacting Figure 6.1.2 Definition of lift and drag
force F from the flow is decomposed into a direction perpendicular to the velocity at infinity V and to
a direction parallel to V . The former component is known as the lift, L; the latter is called the drag, D
(see Figure 6.1.2).
Lift and drag coefficients CL and CD are defined as:
L
CL 
1 2
V c
2  (7.1.1)
and
D
CD  (7.1.2)
1 2
V c
2 
where ρ is the density and c the length of the aerofoil, denoted by the chord. The unit for the lift
and drag in equations (7.1.1) and (7.1.2) is force per length (N/m). A chord line can be defined as the
line from the trailing edge to the nose of the aerofoil (see Figure 6.1.2). To describe the forces
completely it is also necessary to know the moment M about a point in the aerofoil. This point is often
located on the chord line at c/4 from the leading edge. The moment is positive when it tends to turn the
aerofoil in Figure 6.1.2 clockwise (nose up) and a moment coefficient is defined as:
M
CM  (7.1.3)
1 2
V c
2 

| Rotor Blade Theory Pavesi - 124 -


The coefficients CL, CD and CM are functions of
α, Re and Ma.
The two-dimensional steady angle α of attack is
defined as the geometric angle between the
undisturbed stream lines and the chord line of the
profile (Figure 6.1.3). The lift force is by definition
directed perpendicular to the undisturbed inflow
direction. Undisturbed flow is defined as the flow
without the influence of the profile. In two-
dimensional steady flow the changes to the flow field
are only induced by bound vorticity.
Figure 6.1.3 Airfoil characteristics as Re is the Reynolds number based on the chord
function of the angle of attack. and V, Re=cV /ν, where ν is the kinematic viscosity;
and Ma denotes the Mach number, in other words the
ratio between V and the speed of sound. For a wind
turbine and a slow moving aircraft the lift, drag and
moment coefficients are only functions of α and Re.
Lift is described in theory as the force exerted by
a fluid flow on a bound vortex, given that the fluid
flows perpendicular to the vorticity vector. The
vorticity is defined as ω=×V. The total vorticity or
circulation Γ in a surface S is the integral of the local
vorticity over S:
   ds (7.1.4)
  V  dC
Figure 6.1.4 Polar for the FX67-K-170 The equivalence of the integral over the surface S
airfoil and the integral along the closed curve C follows from
Stokes' theorem. The difference between vorticity and
circulation is, that vorticity is a property of an infinitesimal element of fluid, while circulation is an
integral property. The physical meaning of circulation becomes clear when a line (hence a 2d-situation)
of constant vorticity ω is considered. If C is a circle of radius r perpendicular to the line of vorticity in
the centre, it follows that V = Γ/(2πr). The lift force per unit length is related to the circulation and the
inflow velocity via Joukowski’s theorem:
L   V  (7.1.5)
According to Joukowski's hypothesis, the effect of viscosity in the boundary layer is to cause
precisely that circulation so that the stagnation point at the rear of the airfoil corresponds to the sharp
trailing edge of the airfoil. Joukowski's hypothesis implies that the circulation around an airfoil under
small inflow angles is almost proportional to this inflow angle. Airfoils have camber since it yields a
slightly better performance regarding the lift over drag ratio. The camber also causes the lift curve of
the airfoil to shift over a certain angle α0, which is the angle of attack at zero lift (see Figure 6.1.3).
For a given airfoil the behaviours of CL, CD and CM are measured and plotted in so-called polars.
An example of a measured polar for the FX67-K-170 airfoil is shown in Figure 6.1.4.

| Rotor Blade Theory Pavesi - 125 -


Theorycally CL increases linearly with , with a
slope of 2/rad, until a certain value of α, where a
maximum value of CL is reached. In practice the slope
dCL/dα is approximately 5.7 instead of the theoretical 2π.
Typical relations for the lift and drag coefficients as a
function of the angle of attack are presented in Figure
6.1.6. In this figure it can be seen that in practice the lift
curve deviates a great deal from the theoretical curve
beyond approximately α=10°. The reason is that the flow
on the suction side of the airfoil does not reach the
trailing edge any more. At a certain distance from the
trailing edge it comes to a standstill, causing reversal of
the flow and separation. These effects cause a loss of lift
and a sudden increase of drag.
Figure 6.1.6 Reynold influence on the
FX67-K-170 airfoil Lift Coefficient
Hereafter the aerofoil is said to stall and CL
decreases in a very geometrically dependent
manner. For small angles of attack the drag
coefficient CD is almost constant, but increases
rapidly after stall. The Reynolds number
dependency can also be seen in Figure 6.1.6. It is
seen, especially on the drag, that as the Reynolds
number reaches a certain value, the Reynolds
number dependency becomes small. The Reynolds
number dependency is related to the point on the
aerofoil, where the boundary layer transition from
laminar to turbulent flow occurs. The way an
aerofoil stalls is very dependent on the geometry.
Thin aerofoils with a sharp nose, in other words
with high curvature around the leading edge, tend
to stall more abruptly than thick aerofoils.
Different stall behaviours are seen in Figure 6.1.5,
where CL(α) is compared for two different
aerofoils. The FX38-153 is seen to lose its lift
more rapidly than the FX67- K-170.
The explanation lies in the way the boundary
layer separates from the upper side of the aerofoil.
If the separation starts at the trailing edge of the
aerofoil and increases slowly with increasing angle
of attack, a soft stall is observed, but if the
separation starts at the leading edge of the aerofoil,
the entire boundary layer may separate almost
simultaneously with a dramatic loss of lift. The
behaviour of the viscous boundary layer is very Figure 6.1.5 Different stall behaviour

| Rotor Blade Theory Pavesi - 126 -


complex and depends, among other things, on the curvature of the aerofoil, the Reynolds number, the
surface roughness and, for high speeds, also on the Mach number.
One sometimes holds [Hoerner, S.F. and Borst, H.V. (1985)] that 'an airfoil is said to stall when
the lift decreases with increasing angle of attack'. But at angles beyond the stall angle, the lift first
decreases and then increases again and develops a secondary maximum at an angle of attack of
approximately π/4. Moreover under conditions of rotation the airfoil behaviour can change
considerably and the L(α) could even become a monotonous rising function up to an angle of attack of
approximately π/4. And for example in [White, F. M. (1991)] stall is again defined to be equal to
boundary layer separation. This demonstrates that the term 'stall' is not clearly defined. The phenomena
in the flow that cause the loss of lift have a clearer meaning. These phenomena are reversed flow and
separation.
Separation refers to detachment of the boundary layer from the airfoil. The explanation for a two-
dimensional situation is as follows. With increasing angle of attack the circulation increases and the
suction peak near the leading edge becomes deeper. This means that the velocity just outside the
boundary layer near the suction peak becomes very high. The suction peak is located near the
stagnation point where the boundary layer is still very thin. Thus the velocity gradient in the boundary
layer, and thereby the viscous shear stress, becomes very high. This viscous shear converts kinetic
energy from the boundary layer flow into heat. When the flow has passed the suction peak, four
quantities/effects will determine whether it will reach the trailing edge. First, the remained speed
(kinetic energy); second the trajectory of the viscous shear over the surface; third the trajectory of the
adverse pressure gradient which decelerates the flow; and fourth the momentum that will be transferred
from the main stream via viscous and turbulent stress in the boundary layer. The integral of the shear
stress from the stagnation point to a certain position further downstream determines the kinetic energy
losses.
With increasing angle of attack the suction peek becomes deeper since the curvature of the flow
around the leading edge increases. The deeper this suction the more kinetic energy is lost by shear, and
at a certain angle the flow does not reach the trailing edge any more. At a certain position it comes to a
standstill (not only on the surface but also above it; mathematically formulated this means that the
velocity gradient normal to the wall is zero (see (7.1.6)) and this position is called the separation line.
The occurrence of two separation lines is possible in practice (see Figure 6.1.7). Downstream from the
separation line the pressure gradient accelerates the air towards the suction peak and this causes re-
versed flow.
Three types of separation can be distinguished. Two of them concern 'two-dimensional' flow and
the third concerns the rotating case. The 'two-dimensional' types are leading edge separation and
trailing edge separation.
The Leading Edge
Separation has two
appearances: the long
bubble and the short
bubble. The long bubble
type of leading edge
separation and turbulent
reattachment downstream,
is a laminar separation
type and gives a gradual
Figure 6.1.7 Qualitative representation of separation types.

| Rotor Blade Theory Pavesi - 127 -


decrease of the lift curve slope. The bubble grows with
increasing angle of attack towards the trailing edge. It occurs
on thin airfoil sections in combination with low Reynolds
numbers < 5·105. At higher Reynolds number this separation
has less effect although the physical mechanism remains the
same: laminar separation and immediate turbulent reattachment
within the first 1% of the chord. Due to the condition of the low
Reynolds number this separation type is not expected to be
significant on rotors above 5m diameter. A short bubble type
can also be formed on the leading edge, which is quickly
reattached.
Above a certain angle of attack such bubbles suddenly
burst and cause a sudden lift drop and drag increase. This
Figure 6.1.8 Different stall types for occurs on thin airfoil sections with a round nose and low
NACA airfoils at Reynolds 5.5·106 camber, for example the NACA 63-012 (see figure 2.13). Wind
[Hoerner, S.F. and Borst, H.V. turbine blades in general have camber and are thicker, thus they
(1985)]. probably will not suffer from this abrupt type of separation.
The trailing edge type of separation is a gradual type of
separation, which starts at the trailing edge and moves forward with increasing angle of attack. This
occurs on thick or cambered airfoils with a round nose, which have a less deep suction peak, so the
trailing edge becomes the preferential location for the onset of separation. This stall type is usually
observed on airfoil sections for wind turbine rotors in the wind tunnel. Airfoil sections for wind turbine
blades range from 15% thickness to approximately 35% and the Reynolds number is approximately
5·106 for a 1MW rotor of 60m diameter.
The effective shape of an airfoil is the geometrical shape to which the displacement thickness has been
added. This displacement thickness accounts for viscosity effects when the flow around the airfoil is
described by Euler’s potential theory. Important
deviations between the lift characteristics of an airfoil
and those predicted by potential theory occur when
the displacement thickness becomes large, which
corresponds to the occurrence of separation. As
reversed flow is always the consequence of
separation, the occurrence of reversed flow can be
used to measure the onset of significant deviation of
the lift curve slope. Thus the initial occurrence of
reversed flow denotes the onset of large deviations
from the theoretical lift 2π(α-α0). The boundary layer
near the location of separation is often thought to
behave as indicated in van Kuik ( 1991). Here, one Figure 6.1.9 Diagrammatic representation
streamline intersects the wall at the point of of the boundary layer flow near the separation
separation s. The location of s is determined by the point [51].
condition that the velocity grad-.mnbient normal to the wall vanishes there:
 V 
  0 (7.1.6)
 x Wall

| Rotor Blade Theory Pavesi - 128 -


in which x is the distance to the wall. This way of seeing things differs from the description given
by Betz (1935): ‘Very often another phenomenon can be observed in the period of transition between
normal and disturbed (separated) flow, the two states of affairs continually interchanging.’ Recently
Simpson (1996) came up with a more realistic model of turbulent separation. He argued that the
criterion of the vanishing velocity gradient is too narrow for separation and that separation begins
intermittently at a given location. The flow reversal occurs only a fraction of the time.
At progressively downstream locations, the fraction of the time that the flow moves downstream
is progressively less. Quantitative definitions were proposed on the basis of the fraction of the time that
the flow moves downstream. Incipient detachment (ID) is defined as 1% of the time reversed flow,
intermittent transitory detachment (ITD) as 20% of the time reversed flow and transitory detachment
(TD) corresponds to 50% of the time reversed flow.
ID corresponds to the practical situation in which flow markers such as tufts move occasionally in
the reversed direction.

| Rotor Blade Theory Pavesi - 129 -


6.2 Blade Element Theory
The aerodynamic lift (and
drag) forces on the span-wise
elements of radius r and length r
of the several blades of a wind
turbine rotor are responsible for
the rate of change of axial and
angular momentum of all of the
air which passes through the
annulus swept by the blade
elements. In addition, the force on
the blade elements caused by the
drop in pressure associated with
the rotational velocity in the wake
must also be provided by the
aerodynamic lift and drag. As
there is no rotation of the flow
approaching the rotor the reduced Figure 6.2.1 A Blade Element Sweeps Out an Annular Ring
pressure on the downwind side of
the rotor caused by wake rotation
appears as a step pressure drop
just like that which causes the
change in axial momentum.
Because the wake is still rotating
in the far wake the pressure drop
caused by the rotation is still
present and so does not contribute
Figure 6.2.2 Blade Element Velocities and Forces
to the axial momentum change.
It is assumed that the forces
on a blade element can be calculated by means of two dimensional aerofoil characteristics using an
angle of attack determined from the incident resultant velocity in the cross-sectional plane of the
element; the velocity component in the span-wise direction is ignored. Three-dimensional effects are
also ignored.
The velocity components at a radial position on the blade expressed in terms of the wind speed,
the flow factors and the rotational speed of the rotor will determine the angle of attack. Having
information about how the aerofoil characteristic coefficients CL and CD vary with the angle of attack
the forces on the blades for given values of a and a’ can be determined.
Consider a turbine with nb blades of tip radius R each with chord c and set pitch angle β measured
between the aerofoil zero lift line and the plane of the disc. Both the chord length and the pitch angle
may vary along the blade span. Let the blades be rotating at angular velocity  and let the wind speed
be U. The tangential velocity r of the blade element shown in Figure 6.2.1 combined with the
tangential velocity of the wake a’r means that the net tangential flow velocity experienced by the
blade element is (1+a’)r. Figure 6.2.2 shows all the velocities and forces relative to the blade chord
line at radius r.
From Figure 6.2.2 the resultant relative velocity at the blade is

| Rotor Blade Theory Pavesi - 130 -


W  U2 1  a   2 r 2 1  a '
2 2
(7.2.1)

which acts at an angle  to the plane of rotation, such that


U 1  a  U 1  a '
sin   and cos    (7.2.2)
W W
The angle of attack  is then given by
   (7.2.3)
The lift force on a span-wise length r of each blade, normal to the direction of W, is therefore
1
L  W 2cCLr (7.2.4)
2
and the drag force parallel to W is
1
D  W 2cCDr (7.2.5)
2
The basic assumption of the BEM (blade element – momentum) theory is that the force of a blade
element is solely responsible for the change of momentum of the air which passes through the annulus
swept by the element. It is therefore to be assumed that there is no radial interaction between the flows
through contiguous annuli—a condition that is, strictly, only true if the axial flow induction factor does
not vary radially. In practice, the axial flow induction factor is seldom uniform and experimental
examination of flow through propeller discs by Lock (1924) showed that the assumption of radial
independence is acceptable.
The component of aerodynamic force on nb blade elements resolved in the axial direction is
1
Lcos   Dsin   W 2 n bc  CL cos   Cd sin   r (7.2.6)
2
The rate of change of axial momentum of the air passing through the swept annulus is
U 1  a  2rr2aU  4U2 a 1  a  rr (7.2.7)
The drop in wake pressure caused by wake rotation is equal to the increase in dynamic head,
which is
1
  2a ' r 
2
(7.2.8)
2
Therefore the additional axial force on the annulus is
1
  2a ' r  2rr
2
(7.2.9)
2
Thus
1
W 2 n b c  CL cos   CD sin   r  4  U2 a 1  a    a ' r   rr
2
(7.2.10)
2  

| Rotor Blade Theory Pavesi - 131 -


Simplifying,
W2 c
n b  CL cos   CD sin  r  8 a 1  a    a '    r
2
(7.2.11)
2
U R  

The element of axial rotor torque caused by aerodynamic forces on the blade elements is
1
 Lsin   Dcos  r  W 2 n bc  CL sin   CD cos   rr (7.2.12)
2
The rate of change of angular momentum of the air passing through the annulus is
U 1  a  r2a 'r2rr  4U   r  a ' 1  a  r 2r (7.2.13)
Equating the two moments
1
W 2 n b c  CL sin   Cd cos   rr  4U   r  a ' 1  a  r 2r (7.2.14)
2
Simplifying,
W2 c
2
n b  CL sin   Cd cos   82a ' 1  a  (7.2.15)
U R
parameter =r/R.
It is convenient to put
CL cos   CD sin   Cx (7.2.16)
and
CL sin   CD cos   Cy (7.2.17)
Solving equations (7.2.11) and (7.2.15) to obtain values for the flow induction factors a and a’
using two-dimensional aerofoil characteristics requires an iterative process. The following equations,
derived from (7.2.11) and (7.2.15), are convenient in which the right-hand sides are evaluated using
existing values of the flow induction factors yielding simple equations for the next iteration of the flow
induction factors.
a r  r 
2  x 
 C 
2
C2y  (7.2.18)
1  a 4sin   4sin  
2

a' r C y
 (7.2.19)
1  a ' 4sin  cos 
Blade solidity  is defined as total blade area divided by the rotor disc area and is a primary
parameter in determining rotor performance. Chord solidity r is defined as the total blade chord length
at a given radius divided by the circumferential length at that radius.
nb c nb c
r   (7.2.20)
2 r 2 R

| Rotor Blade Theory Pavesi - 132 -


It is argued by Wilson and Lissaman (1974) that the drag coefficient should not be included in
Equations (7.2.18) and (7.2.19) because the velocity deficit caused by drag is confined to the narrow
wake which flows from the trailing edge of the aerofoil.
Furthermore, Wilson and Lissaman reason, the drag-based velocity deficit is only a feature of the
wake and does not contribute to the velocity deficit upstream of the rotor disc. The basis of the
argument for excluding drag in the determination of the flow induction factors is that, for attached
flow, drag is caused only by skin friction and does not affect the pressure drop across the rotor. Clearly,
in stalled flow the drag is overwhelmingly caused by pressure. In attached flow it has been shown by
Young and Squire (1938) that the modification to the inviscid pressure distribution around an aerofoil
caused by the boundary layer has an affect both on lift and drag.
The ratio of pressure drag to total drag at zero angle of attack is approximately the same as the
thickness to chord ratio of the aerofoil and increases as the angle of attack increases. One last point
about the BEM theory: the theory is strictly only applicable if the blades have uniform circulation, i.e.,
if a is uniform. For non-uniform circulation there is a radial interaction and exchange of momentum
between flows through adjacent elemental annular rings. It cannot be stated that the only axial force
acting on the flow through a given annular ring is that due to the pressure drop across the disc.
However, in practice, it appears that the error involved in relaxing the above constraint is small for tip
speed ratios greater than 3.

6.3 Determination of rotor torque and power


The calculation of torque and power developed by a rotor requires a knowledge of the flow
induction factors, which are obtained by solving equations (7.2.18) and (7.2.19). The solution is usually
carried out iteratively because the two-dimensional aerofoil characteristics are non-linear functions of
the angle of attack.
To determine the complete performance characteristic of a rotor, i.e., the manner in which the
power coefficient varies over a wide range of tip speed ratio, requires the iterative solution. The
iterative procedure is to assume a and a’ to be zero initially, determining , Cp and Cd on that basis, and
then to calculate new values of the flow factors using Equations (7.2.18) and (7.2.19). The iteration is
repeated until convergence is achieved.
From Equation (7.2.14) the torque developed by the blade elements of span-wise length r is
M  4U  r  a ' 1  a  r 2r (7.3.1)
If drag, or part of the drag, has been excluded from the determination of the flow induction
factors then its effect must be introduced when the torque caused by drag is calculated from blade
element forces, see equation (7.2.12),
1
M  4U  r  a ' 1  a  r 2r  W 2n bcCd cos    rr (7.3.2)
2
The complete rotor, therefore, develops a total torque M:
  c  
n
1 2 3  2   
R
W bR
M  U R    8a ' 1  a    Cd 1  a '  d  (7.3.3)
2 0  U   
   

| Rotor Blade Theory Pavesi - 133 -


The power developed by the rotor is P=M. The power coefficient is
P
CP  (7.3.4)
1 3 2
U  R
2
Solving the BEM Equations (7.2.18) and (7.2.19) for a given, suitable blade geometrical and
aerodynamic design yields a series of values for the power and torque coefficients which are functions
of the tip speed ratio. A typical performance curve for a modern, high-speed wind turbine is shown in
Figure 6.3.1.
The maximum power coefficient occurs at
a tip speed ratio for which the axial flow
induction factor a, which in general varies with
radius, approximates most closely to the Betz
limit value of 1/3. At lower tip speed ratios the
axial flow induction factor can be much less
than 1/3 and aerofoil angles of attack are high
leading to stalled conditions. For most wind
turbines stalling is much more likely to occur at
the blade root because, from practical
constraints, the built-in pitch angle β of a blade
is not large enough in that region. At low tip
speed ratios blade stalling is the cause of a
significant loss of power, as demonstrated in
Figure 6.3.1. At high tip speed ratios a is high,
angles of attack are low and drag begins to Figure 6.3.1 Power Coefficient-Tip Speed Ratio
predominate. At both high and low tip speed Performance Curve
ratios, therefore, drag is high and the general
level of a is non-optimum so the power coefficient is low. Clearly, it would be best if a turbine can be
operated at all wind speeds at a tip speed ratio close to that which gives the maximum power
coefficient.

Figure 6.3.2 Typical variations of Power Figure 6.3.3 Typical variations of Torque Coeffi
Coefficient for a variable-pitch wind cient for a variable-pitch wind
turbine turbine

| Rotor Blade Theory Pavesi - 134 -


The power coefficient CP has its maximum
at (λo, βo), with βo being a very small angle,
ideally zero. This has a pair of important
connotations. On the one hand, the maximum at
β = 0 means that any deviation of the pitch angle
yields lower power capture. On the other hand,
maximum conversion efficiency is
accomplished at λo. So, fixed-speed turbines
will operate with maximum efficiency just for a
unique wind speed, whereas variable-speed
turbines can potentially work with maximum
efficiency over a wide wind speed range at least
up to rated power. To realise the potential
benefits of variable-speed operation, the Figure 6.3.4 Typical variations of C and C for a
M P
rotational speed must be adjusted initially in fixed-pitch wind turbine.
proportion to the wind speed to maintain an
optimum tip-speed-ratio. It can also be observed that the maximum of CM occurs at (λQmax, βo), with
λMmax < λo.

| Rotor Blade Theory Pavesi - 135 -


7 Breakdown of the Momentum Theory
7.1 Free-stream/wake mixing
For heavily loaded turbines, when a is high, the momentum theory predicts a reversal of the flow
in the wake. Such a situation cannot actually occur so what happens is that the wake becomes turbulent
and, in doing so, entrains air from outside the wake by a mixing process which re-energizes the slow
moving air which has passed through the rotor.
A rotor operating at increasingly high tip speed ratios presents a decreasingly permeable disc to
the flow. Eventually, when  is high enough for the axial flow factor to be equal to one, the disc
effectively becomes a solid plate.
The flow past a solid disc, because of viscosity, separates at the disc’s edge. A boundary layer
develops as the flow over the front of the disc spreads out radially and by the time it reaches the edge
viscosity has sapped much of the kinetic energy.
As the boundary layer flows around the disc edge it accelerates causing a large drop in static
pressure (Bernoulli). To flow around the disc edge would require very high velocity and there is
insufficient static pressure to provide the necessary kinetic energy. The flow, therefore, separates from
the disc and continues in the general stream-wise direction. In the region directly behind the disc there
is slow moving, almost stagnant, air at the low static pressure of the flow separating at the disc edge. At
the front of the disc, at the very centre, the flow is brought to rest and so there is a large increase in
static pressure as the kinetic energy is converted to pressure energy. Elsewhere on the front surface the
flow moves radially with a velocity, outside the boundary layer, which increases towards the disc edge.
The static pressure is generally higher on the front of the disc than on the rear and so the disc
experiences a pressure drag force.
A similar process happens with a spinning rotor at high tip speed ratios. The air which does not
pass through the rotor disc moves radially outwards and separates at the disc edge causing a low static
pressure to develop behind the disc; the drop in static pressure caused by the separation increases as the
tip speed ratio rises and the axial flow factor increases. The air which does pass through the rotor
emerges into a low pressure region and is moving slowly. There is insufficient kinetic energy to
provide the rise in static pressure necessary to achieve the ambient atmospheric pressure that must exist
in the far wake. The air can only achieve atmospheric pressure by gaining energy from the mixing
process in the turbulent wake. The shear layer in the flow between the free-stream air and the wake air
is what becomes of the boundary layer that develops on the front of the disc. The shear layer is unstable
and breaks up into the turbulence that causes the mixing and reenergization of the wake air.

7.2 Modification of rotor thrust caused by flow separation


The low static pressure downstream of the rotor disc caused by the separation of the free-stream
flow at the edge of the disc and the high static pressure at the stagnation point on the upstream side
causes a large thrust on the disc, much larger than that predicted by the momentum theory. Some
experimental results reported by Glauert (1926) for a whole rotor can be seen in Figure 7.4.1 where the
simple expression for the thrust force coefficient, as derived from the momentum theory (CT=4a(1- a)),
is given for comparison.
The thrust (or drag) coefficient for a simple, flat circular plate is given by Hoerner (1965) as 1.17
but, as demonstrated in Figure 7.2.1, the thrust on the rotating disc is higher. It might have been
expected that when a=1 the rotor would have the same thrust coefficient as the circular plate. The

| Breakdown of the Momentum Theory Pavesi - 136 -


principal difference between the
circular plate and the rotor is that the
latter is rotating and, as Hoerner also
describes, this causes energy to be
dissipated in a thicker, rotating
boundary layer on the upstream
surface of the rotor disc giving rise
to an even lower pressure on the
downstream side.
It would follow from the
above arguments that for high
values of the axial induction factor
most of the pressure drop across the
disc is not associated with blade
circulation, there is no circulation
Figure 7.2.1 Comparison of Theoretical and Measured Values present in the case of the circular
of CT plate.
Circulation would cause a
small pressure drop similar to that given by the momentum theory because it would be determined by
the very low axial velocity of the flow which actually permeates the disc.

7.3 Empirical determination of thrust coefficient


A suitable straight line through the experimental points would appear to be possible, although
Glauert proposed a parabolic curve, and provides an empirical solution to the problem of the thrust on a
heavily loaded turbine (a rotor operating at a high value of the axial flow induction factor).
Most authors assume that the entire thrust on the rotor disc causes axial momentum change.
Therefore, for the empirical line to be useful it must be assumed that it applies not only to the whole
rotor but also to each separate streamtube.
Let CT1 be the empirical value of CT when a=1. Then, as the straight line must be a tangent to the
momentum theory parabola at the transition point, the equation for the line is
CT  CT1  4  
CT1  1 1  a  (8.3.1)

and the value of a at the transition point is


1
aT  1 CT1 (8.3.2)
2
By inspection of Figure 7.2.1, CT1 must lie between 1.6 and 2; CT1=1.816 would appear to be the
best fit to the experimental data of Figure 7.2.1, whereas Wilson and Lissaman (1974) favour the lower
value of CT1=1.6. Glauert fits a parabolic curve to the data giving much higher values of C T1 at high
values of a but he was considering the case of an airscrew in the windmill brake state where the angles
of attack are negative.
The flow field through the turbine under heavily loaded conditions cannot be modelled easily and
the results of this empirical analysis must be regarded as being only approximate at best. They are,
nevertheless, better than those predicted by the momentum theory. For most practical designs the value

| Breakdown of the Momentum Theory Pavesi - 137 -


of axial flow induction factor rarely exceeds 0.6 and for a well-designed blade will be in the vicinity of
0.33 for much of its operational range.
For values of a greater than aT it is common to replace the momentum theory thrust in equation
(6.1.22) with equation (8.3.1), in which case Equation (7.2.18) is replaced by

1  a 
2 r
sin 2
Cx  4  
CT1  1 1  a   CT1  0 (8.3.3)

in which the pressure drop caused by wake rotation in ignored as it is very small.
However, as the additional pressure drop is caused by edge flow separation then this course of
action is questionable and it may be more appropriate to retain equation (7.2.18).

7.4 The Effects of a Discrete Number of Blades


The analysis described in the previous sections of this chapter assumes that there is a sufficient
number of blades on the rotor for every fluid particle passing through the rotor disc to interact with a
blade, i.e., that all fluid particles undergo the same loss of momentum. With a small number of blades
some fluid particles will interact with them but most will pass between the blades and, clearly, the loss
of momentum by a particle will depend on its proximity to a blade as the particle passes through the
rotor disc. The axial induced velocity will, therefore, at any instant, vary around the disc, the average
value determining the overall axial momentum of the flow and the larger value local to a blade
determining the forces on the blade.

7.4.1 Tip-losses
If the axial flow induction factor a is large at the blade position then, by equation (7.2.2), the
inflow angle  will be small and the lift force will be almost normal to the rotor plane. The component
of the lift force in the tangential direction
will be small and so will be its
contribution to the torque. A reduced
torque means reduced power and this
reduction is known as tip loss because the
effect occurs only at the outermost parts
of the blades.
In order to account for tip losses, the
manner in which the axial flow induction
factor varies azimuthally needs to be
known but, unfortunately, this
requirement is beyond the abilities of the
blade element–momentum theory.
Just as a vortex trails from the tip of
an aircraft wing so does a vortex trail
from the tip of a wind turbine blade. Figure 7.4.1 Helical Trailing Tip Vortices of a Horizontal
Because the blade tip follows a circular Axis Turbine Wake
path it leaves a trailing vortex of a helical
structure, as is shown in Figure 7.4.1, which convects downstream with the wake velocity. For a two-
blade rotor, unlike an aircrafts wings, the bound circulations on the two blades are opposite in sign and

| Breakdown of the Momentum Theory Pavesi - 138 -


so combine to shed a straight
line vortex along the rotational
axis with strength equal to the
blade circulation times the
number of blades.
For a single vortex to be
shed from the blade tip the
circulation strength along the
blade span must be uniform and,
as has been shown, uniform
circulation is a requirement for
optimized operation. However,
the uniform circulation
requirement assumes that the
axial flow induction factor is
uniform across the disc and, as
has been argued above, with Figure 7.4.2 Azithimuthal Variation of a for Various Radial
discrete blades rather than a Positions for a Three-blade Rotor with Uniform Blade Circulation
uniform disc the flow factor is Operating at a Tip Speed Ratio of 6. The blades are at 120°, 240°
not uniform. and 360°.
In the case of Figure 7.4.1,
very close to the blade tips the
tip vortex causes very high
values of the flow factor a such
that, locally, the net flow past
the blade is in the upstream
direction. The average value of
a, azimuthally, is radially
uniform which means that if
high values occur in the vicinity
of the blades then low values
Figure 7.4.3 Span-wise Variation of the Tip-loss Factor for a
occur elsewhere. The azimuthal
Blade with Uniform Circulation
variation of a at various radial
positions is shown in Figure 7.4.2 for a three blade rotor operating at a top speed ratio of 6. The
calculation for Figure 7.4.3 assumes a discrete vortex for each blade with a constant pitch and constant
radius helical wake.
The ratio of the average value of a to that at a blade position is shown in Figure 7.4.2 and only
near the tip does the ratio begin to fall to zero. The ratio is called ‘the tip-loss factor’.
From Equation (6.3.10) and in the absence of tip-loss and drag the contribution of each blade
element to the overall power coefficient is

CP  8 23a ' 1  a   (8.4.1)


Substituting for a’ from Equation (6.3.14) gives

| Breakdown of the Momentum Theory Pavesi - 139 -


CP  8a 1  a  
2
(8.4.2)

Whereas, from the Kutta–Joukowski theorem, the circulation  on the blade, which is uniform,
provides a torque per unit span of
M
   W    sin r r (8.4.3)
r
where the angle  is determined by the flow velocity local to the blade.
The strength of the total circulation for all blades is given by
U2
  4 a 1  a  (8.4.4)

and so, in the presence of tip-loss the increment of power coefficient from a blade element is
CP  8a 1  a 1  a r   (8.4.5)
where a=1/3 is the average axial flow induction factor and ar is the value local to the blade.
The results from Equations (8.4.4) and (8.4.5) are plotted in Figure 7.4.3 and clearly show the
effect of tip-loss. Equation (8.4.2) assumes that a=1/3 uniformly over the whole disc, Equation (8.4.5)
recognizes that a is not uniform. The azimuthally averaged value of a is equal to 1/3 at every radial
position but the azimuth variation gives rise to the tip-loss. The blade does not extract energy from the
flow efficiently because a varies. Imagine the disc comprising a myriad of elemental discs, each with
its own independent stream-tube, and not all of them operating at the Betz limit. Note that the power
loss to the wind is exactly the same as that extracted by the blades, there is no effective drag associated
with tip-loss.
With uniform circulation the azimuthal average value of a is also radially uniform but that implies
a discontinuity of axial velocity at the wake boundary with a corresponding discontinuity in pressure.
Whereas such discontinuities are acceptable in the idealized actuator disc situation they will not occur
in practice with a finite number of blades. If it is assumed that a is zero outside of the wake then a must
fall to zero in a regular fashion towards the blade tips and, consequently, the bound circulation must
also fall to zero. The manner in which the circulation varies at the tip will be governed by the blade tip
design, that is, the chord and pitch variation, and there will be a certain design which will minimize the
tip-loss.
If the circulation varies along the blade span vorticity is shed into the wake in a continuous
fashion from the trailing edge. Therefore, each blade sheds a helicoidal sheet of vorticity, as shown in
Figure 7.4.5, rather than a single helical vortex as shown in Figure 7.4.2. The helicoidal sheets convect
with the wake velocity and so there can be no flow across the sheets which can therefore be regarded as
impermeable.
The intensity of the vortex sheets is equal to the rate of change of bound circulation along the
blade span and so increases rapidly towards the blade tips. There is flow around the blade tips because
of the pressure difference between the blade surfaces which means that on the upwind surface of the
blades the flow moves towards the tips and on the downwind surface the flow moves towards the root.
The flows from either surface leaving the trailing edge of a blade will not be parallel to one another and
will form a surface of discontinuity of velocity in a radial sense within the wake; the axial velocity
components will be equal. The surface of discontinuity is called a vortex sheet. A similar phenomenon

| Breakdown of the Momentum Theory Pavesi - 140 -


occurs with aircraft wings
and a textbook of aircraft
aerodynamics will explain it
in greater detail.
A deeper understanding
of the mechanism of tip-loss
can be obtained by following
the path of air particles. An
air particle approaches the
spinning rotor, ‘senses’ high
pressure ahead and slows
down accordingly. The high
pressure on the upwind side
of the rotor blades is
effectively smeared around
the whole disc.
Slowing down also
causes the particle to move Figure 7.4.4 Span-wise Variation of Power Extraction in the Presence
outwards to maintain the of Tip-loss for a Blade with Uniform Circulation on a Three-blade
mass flow rate. When the Turbine Operating at a Tip Speed Ratio of 6
particle reaches the rotor
plane it will either be close to
a blade or not and its axial
velocity will be affected
accordingly, as shown in
Figure 7.4.2. If the particle
passes through the rotor
plane close to blade then it
will also be strongly affected
by the blade’s pressure field.
A particle which passes
close to and in front of a
blade will leave the trailing
edge having accelerated in
the tangential direction; it
will then pass downstream,
on the upwind side of the Figure 7.4.5 A Helicoidal Vortex Sheet Wake for a Two-blade Rotor
vortex sheet being shed from the trailing edge and so will also be moving radially outwards. The
particle, therefore, migrates outward to the edge of the vortex sheet around which it is swept on to the
downwind side and migrates inward with a radial velocity which reduces to zero at a radial point on the
sheet where the shed vorticity is zero. The particle then continues downstream with the velocity of the
axial and tangential velocities of the vortex sheet.
A second particle which passes a blade close to the downwind, low pressure, surface is
accelerated tangentially in the opposite direction to the blade motion and then slows down, leaving the
trailing edge with the same axial and tangential velocity components as the first particle but on the

| Breakdown of the Momentum Theory Pavesi - 141 -


downwind side of the vortex sheet so it will have, in addition, a radially inwards velocity. The second
particle will, depending on its radial position, migrate inwards until the radial velocity becomes zero.
A third particle which passes between two blades will be moving axially at a greater velocity than
the first two particles, will not be strongly affected by the pressure fields of the blades but, because of
the solid blockage presented by the blades Figure 7.4.1, will be directed into a helical path. Being
faster, axially, than the vortex sheet ahead the particle will begin to catch up, as it does so the influence
of the vortex sheet will move it outwards, around the edge of the sheet and then inwards, just like the
first particle. Unlike the first particle, however, the third particle will still be moving faster than the
vortex sheet and so will move axially away from the sheet, approaching the next sheet downstream and
repeating the motion around the edge of that sheet. The particle will proceed downstream overtaking
and hopping around each vortex sheet in turn.
The third particle does not lose as much axial momentum as particles one and two and is therefore
affected by the so-called tip-loss. The affect is greater the closer the third particle is to the edge of the
rotor disc as it passes through the disc.
A fourth particle passes between the blades but at a radial position, closer to the axis of rotation,
where its axial velocity is equal to that of the vortex sheets. If the particle passes midway, say, between
two blades then it remains midway between the two corresponding vortex sheets as it moves
downstream and does not undergo any radial motion other than the general expansion caused by the
slowing down of the flow. The fourth particle is totally unaffected by the fact that there is a finite
number of blades and follows the same progress as if it were passing through a uniform actuator disc.
The axial flow induction factor varies, therefore, not only azimuthally but also radially, is a
function of both r and . The azimuthally averaged value of a(r)=ab(r) f (r), where f (r) is known as the
tip-loss factor, has a value of unity inboard (particle 4) and falls to zero at the edge of the rotor disc.
The value ab(r) is the level of axial flow induction factor that occurs locally at a blade element and is
the velocity with which the vortex sheet convects downstream. If ab(r) can be held radially uniform
then the vortex sheets will be radially flat, as shown in Figure 7.4.4, but if ab(r) is not uniform the
vortex sheets will warp.
In the application of the blade element–momentum theory it is argued that the rate of change of
axial momentum is determined by the azimuthally averaged value of axial flow induction factor,
whereas the blade forces are determined by the value of the flow factor which occurs locally at the
blade element, that experienced by the first and second particles.
The mass flow rate through an annulus  U 1  a b  r  f  r  2rr
The azimuthally averaged overall change of axial velocity  2a b  r  f  r  U
The rate of change of axial momentum  4rU2 1  a b  r  f  r  a b  r  f  r  r
1 1
The blade element forces  W 2 n b CL and  W 2 n b CD
2 2
where W, CL and CD are determined using ab(r).
The pressure force caused by the rotation of the wake is also calculated using an azimuthally
averaged value of the tangential flow induction factor 2a 'b  r  f  f  .

| Breakdown of the Momentum Theory Pavesi - 142 -


8 Wind –Turbine Performance
The performance of a wind turbine can be characterized by the manner in which the three main
indicators—power, torque and thrust—vary with wind speed. The power determines the amount of
energy captured by the rotor, the torque developed determines the size of the gear box and must be
matched by whatever generator is being driven by the rotor. The rotor thrust has great influence on the
structural design of the tower. It is usually convenient to express the performance by means of non-
dimensional, characteristic performance curves from which the actual performance can be determined
regardless of how the turbine is operated, e.g., at constant rotational speed or some regime of variable
rotor speed. Assuming that the aerodynamic performance of the rotor blades does not deteriorate the
non-dimensional aerodynamic performance of the rotor will depend upon the tip speed ratio and, if
appropriate, the pitch setting of the blades. It is usual, therefore, to display the power, torque and thrust
coefficients as functions of tip speed ratio.

8.1 The CP –  performance curve


The theory described in Chapter 6.2 gives the wind turbine designer a means of examining how
the power developed by a turbine is governed by the various design parameters. The usual method of
presenting power performance is the adimensional CP –  curve and the curve for a typical, modern,
three-blade turbine is shown in Figure 8.1.1 CP –  Performance Curve for a Modern Three-blade
Turbine Showing Losses..
The first point to notice is that the maximum value of CP is only 0.47, achieved at a tip speed ratio
of 7, which is much less than the Betz limit. The discrepancy is caused, in this case, by drag and tip
losses but the stall also reduces the CP at low values of the tip speed ratio (Figure 8.1.1).
Even with no losses included in the analysis the Betz limit is not reached because the blade design
is not perfect.

8.1.1 The effect of solidity on performance


At this stage, the other principal parameter to consider is the solidity, defined as total blade area
divided by the swept
area. For the three-
blade machine above
the solidity is 0.0345
but this can be altered
readily by changing its
number of blades.
The solidity
could also have been
changed by changing
the blade chord.
The main effects
to observe of changing
solidity are as follows,
see Figure 8.1.2.
Figure 8.1.1 CP –  Performance Curve for a Modern Three-blade Turbine (1) Low solidity
Showing Losses.

| Wind –Turbine Performance Pavesi - 143 -


produces a broad, flat
curve which means
that the CP will
change very little
over a wide tip speed
ratio range but the
maximum CP is low
because the drag
losses are high (drag
losses are roughly
proportional to the
cube of the tip speed
ratio).
(2) High
solidity produces a
narrow performance
curve with a sharp
peak making the
Figure 8.1.2 Effect of Changing Solidity turbine very sensitive
to tip speed ratio
changes and, if the solidity is too high, has a relatively low maximum CP. The reduction in CPmax is
caused by stall losses.
(3) An optimum solidity appears to be achieved with three blades, but two blades might be an
acceptable alternative because although the maximum CP is a little lower the spread of the peak is
wider and that might result in a larger energy capture.
It might be argued that a good solution would be to have a large number of blades of small
individual solidity but
this greatly increases
production costs and
results in blades which
are structurally weak
and very flexible.
There are
applications which
require turbines of
relatively high
solidity, one is the
directly driven water
pump and the other is
the very small turbine
used for battery
charging. In both cases
it is the high starting
torque (high torque at
very low tip speed
Figure 8.1.3 The Effect of Solidity on Torque

| Wind –Turbine Performance Pavesi - 144 -


ratios) which is of importance and this also allows small amounts of power to be developed at very low
wind speeds, ideal for trickle charging batteries.

8.1.2 The CM –  curve


The torque coefficient is derived from the power coefficient simply by dividing by the tip speed
ratio and so it does not give any additional information about the turbine’s performance. The principal
use of the CM– curve is for torque assessment purposes when the rotor is connected to a gear box and
generator.
Figure 8.1.3 shows how the torque developed by a turbine rises with increasing solidity. For
modern high-speed turbines designed for electricity generation as low a torque as possible is desirable
in order to reduce gearbox costs. On the other hand the multi-bladed, high-solidity turbine, developed
in the nineteenth century for water pumping, rotates slowly and has a very high starting torque
coefficient necessary for overcoming the torque required to start a positive displacement pump.
The peak of the torque curve occurs at a lower tip speed ratio than the peak of the power curve.
For the highest solidity shown in Figure 8.1.3 the peak of the curve occurs while the blade is stalled.

8.1.3 The CT –  curve


The thrust force
on the rotor is directly
applied to the tower
on which the rotor is
supported and so
considerably
influences the
structural design of
the tower. Generally,
the thrust on the rotor
increases with
increasing solidity
(Figure 8.1.4).

Figure 8.1.4 The Effect of Solidity on Thrust

| Wind –Turbine Performance Pavesi - 145 -


8.2 Constant Rotational Speed Operation
The majority of wind turbines currently installed generate electricity. Whether or not these
turbines are grid connected they need to produce an electricity supply which is of constant frequency or
else many common appliances will not function properly. Consequently, the most common mode of
operation for a wind turbine is constant rotational speed. Connected to the grid a constant speed turbine
is automatically controlled whereas a stand-alone machine needs to have speed control and a means of
dumping excess power.

8.2.1 The KP – 1/ curve


An alternative performance curve can be produced for a turbine controlled at constant speed. The
CP –  curve shows, non-dimensionally, how the power would vary with rotational speed if the wind
speed was held constant. The KP – 1/ curve describes, again non-dimensionally, how the power
would change with wind speed when constant rotational speed is enforced. KP is defined as
Power CP
Kp   (9.2.1)
1
  R  A d
3 3
2
The CP –  and KP – 1/ curves for a typical fixed-pitch wind turbine are shown in Figure 8.2.1
The KP – 1/ curve, as stated above, has the same form as the power – wind speed characteristic of the
turbine. The efficiency of the turbine (given by the CP –  curve) varies greatly with wind speed, a
disadvantage of constant speed operation, but it should be designed such that the maximum efficiencies
are achieved at wind speeds where there is the most energy available.

8.2.2 Stall regulation


An important feature of this KP – 1/ curve is that the power, initially, falls off once stall has
occurred and then gradually increases with wind speed. This feature provides an element of passive
power output regulation, ensuring that the generator is not overloaded as the wind speed increases.
Ideally, the power should rise with wind speed to the maximum value and then remain constant

Figure 8.2.1 Non-dimensional Performance Curves for Constant Speed Operation

| Wind –Turbine Performance Pavesi - 146 -


regardless of the increase in wind speed; this is called perfect stall regulation. However, stall regulated
turbines do not exhibit the ideal, passive stall behaviour.
Stall regulation provides the simplest means of controlling the maximum power generated by a
turbine to suit the sizes of the installed generator and gearbox and until recently, at the time of writing,
is the most commonly adopted control method.
The principal advantage of stall control is simplicity but there are significant disadvantages. The
power versus wind speed curve is fixed by the aerodynamic characteristics of the blades, in particular
the stalling behaviour. The post stall power output of a turbine varies very unsteadily and in a manner
which, so far, defies prediction. The stalled blade also exhibits low vibration damping because the flow
about the blade is unattached to the low pressure surface and blade vibration velocity has little effect on
the aerodynamic forces. The low damping can give rise to large vibration displacement amplitudes
which will inevitably be accompanied by large bending moments and stresses, causing fatigue damage.
When parked in high, turbulent winds a rotor with fixed pitch blades may well be subject to large
aerodynamic loads which cannot be alleviated by adjusting (feathering) the blade pitch angle.
Consequently, a fixed pitch, stall-regulated turbine
experiences more severe blade and tower loads
than a pitch regulated turbine.

8.2.3 Effect of rotational speed change


The power output of a turbine running at
constant speed is strongly governed by the chosen,
operational rotational speed. If a low rotation
speed is used the power reaches a maximum at a
low wind speed and consequently it is very low.
To extract energy at wind speeds higher than the
stall peak the turbine must operate in a stalled
condition and so is very inefficient. Conversely, a
turbine operating at a high speed will extract a
great deal of power at high wind speeds but at Figure 8.2.2 Effect on Extracted Power of
moderate wind speeds it will be operating Rotational Speed
inefficiently because of the high drag losses.
Figure 8.2.2 demonstrates the sensitivity to
rotation speed of the power output – a 33 percent
increase in r.p.m. from 45 to 60 results in a 150
percent increase in peak power, reflecting the
increased wind speed at which peak power occurs
at 60 r.p.m.
At low wind speeds, on the other hand there
is a marked fall in power with increasing
rotational speed as shown in Figure 8.2.3. In fact,
the higher power available at low wind speeds if a
lower rotational speed is adopted has led to two
speed turbines being built. Operating at one fixed
speed which maximizes energy capture at wind
Figure 8.2.3 Effect on Extracted Power of
speeds at, or above, the average level will result in
Rotational Speed at Low Wind Speed

| Wind –Turbine Performance Pavesi - 147 -


a rather high cut-in wind speed, the lowest wind speed at which generation is possible. Employing a
lower rotational speed at low wind speeds reduces the cut-in wind speed and increases energy capture.
The increased energy capture is, of course, offset by the cost of the extra machinery.

8.2.4 Effect of blade pitch angle change


Another parameter which affects the power output is the pitch setting angle of the blades. Blade
designs almost always involve twist but the blade can be set at the root with an overall pitch angle. The
effects of a few degrees of pitch are shown in Figure 8.2.4.
Small changes in pitch setting angle can
have a dramatic effect on the power output.
Positive pitch angle settings increase the design
pitch angle and so decrease the angle of
incidence. Conversely, negative pitch angle
settings increase the angle of incidence and may
cause stalling to occur as shown in Figure 8.2.4.
A turbine rotor designed to operate optimally at a
given set of wind conditions can be suited to other
conditions by appropriate adjustments of blade
pitch angle and rotational speed.

8.2.5 Pitch regulation


Many of the shortcomings of fixed
pitch/passive stall regulation can be overcome by
providing active pitch angle control Figure 8.2.4.
Figure 8.2.4 Effect on Extracted Power of Blade
shows the sensitivity of power output to pitch
Pitch Set Angle
angle changes.
The most important application of pitch
control is for power regulation but pitch control
has other advantages. By adopting a large positive
pitch angle a large Electrical power (kW) starting
torque can be generated as a rotor begins to turn.
A 90° pitch angle is usually used when shutting
down because this minimises the rotor idling
speed at which the parking brake is applied. At
90° of positive pitch the blade is said to be
‘feathered’. The principal disadvantages of pitch
control are reliability and cost.
Power regulation can be achieved either by
pitching to promote stalling or pitching to feather
which reduces the lift force on the blades by
reducing the angle of attack. Figure 8.2.5 Pitching to Feather Power Regulation
Requires Large Changes of Pitch Angle
8.2.6 Pitching to stall
Figure 8.2.5shows the power curves for a turbine rated at 60 kW, which is achieved at 12 m/s. At

| Wind –Turbine Performance Pavesi - 148 -


wind speeds below the rated level the blade pitch angle is kept at 0°. As rated power is reached only a
small negative pitch angle, initially of about 2°, is necessary to promote stalling and so to limit the
power to the rated level. As the wind speed increases small adjustments in both the positive and
negative directions are all that are needed to maintain constant power. The small sizes of the pitch
angle adjustments make pitching to stall very attractive to designers but the blades have the same
damping and fatigue problems as fixed pitch turbines.

8.2.7 Pitching to feather


By increasing the pitch angle as rated power is reached the angle of attack can be reduced. A
reduced angle of attack will reduce the lift force and the torque. The flow around the blade remains
attached. Figure 8.2.5 is for the same turbine as Figure 8.2.4 but only the zero degree power curve is
shown below the rated level.
Above the rated level fragments of power curves for higher pitch angles are shown as they cross
the rated power line; the crossing points give the necessary pitch angles to maintain rated power at the
corresponding wind speeds. As can be seen in Figure 8.2.5, the required pitch angles increase
progressively with wind speed and are generally much larger than is needed for the pitching to stall
method. In gusty conditions large pitch excursions are needed to maintain constant power and the
inertia of the blades will limit the speed of the control system’s response.
Because the blades remain unstalled if large gusts occur at wind speeds above the rated level
large changes of angle of attack will take place with associated large changes in lift. Gust loads on the
blades can therefore be more severe than for stalled blades. The advantages of the pitching to feather
method are that the flow around the blade remains attached, and so well-understood, and provides
good, positive damping. Feathered blade parking and assisted starting are also available.
Pitching to feather has been the preferred pitch control option mainly because the blade loads can
be predicted with more confidence than for stalled blades.

8.3 Variable-speed Operation


If the speed of the rotor can be continuously adjusted such that the tip speed ratio remains
constant at the level which gives the maximum CP then the efficiency of the turbine will be
significantly increased. Active pitch control is necessary to maintain a constant tip speed ratio but only
in the process of adjustment of rotational speed; the pitch angle should always return to the optimum
setting for highest efficiency. Pitch control regulation is also required in conditions above the rated
wind speed when the rotational speed is kept constant.

8.4 Reference
Abbot, H. and von Doenhoff, A. E. (1959) Theory of Wing Sections, Dover Publications, New York
Batchelor, G.K., An Introduction to Fluid Dynamics', Cambridge University Press, ISBN 0 521
098173, 1967.
Betz, A., (1919). Schraubenpropeller mit geringstem Energieverlust. Gottinger Nachr., Germany.
Betz, A. 'Applied Airfoil Theory' Volume IV, Division J in Durand, W.F. 'Aerodynamic Theory, A
General Review of Progress', 1935.
Coleman, R. P., Feingold, A. M. and Stempin, C. W., (1945). ‘Evaluation of the induced velocity field
of an idealised helicopter rotor’. N.A.C.A. A.R.R,. No. L5E10.

| Wind –Turbine Performance Pavesi - 149 -


de Vries, O. (1979) Fluid Dynamic Aspects of Wind Energy Conversion, AGARDograph No 243,
Advisory Group for Aeronautical Research and Development
Fuglsang, P. and Bak, C. (2003) ‘Status of the Riso wind turbine aerofoils’, presented at the European
Wind Energy Conference, EWEA, Madrid, 16–19 June
Gilbert, B. L. and Foreman, K. M. (1983) ‘Experiments with a diffuser-augmented model wind
turbine’, Journal of Energy Resources Technology, vol 105, pp46–53
Glauert, H., (1926a). ‘The analysis of experimental results in the windmill brake and vortex ring states
of an airscrew’. ARCR R&M No. 1026.
Glauert, H., (1926b). ‘A general theory of the autogyro’. ARCR R&M No. 1111.
Glauert, H., (1935a). ‘Airplane propellers’. Aerodynamic theory (ed. Durand, W. F.). Julius Springer,
Berlin, Germany.
Glauert, H., (1935b). ‘Windmills and fans’. Aerodynamic theory (ed. Durand, W. F.). Julius Springer,
Berlin, Germany.
Goankar, G. H. and Peters, D. A., (1988). ‘Review of dynamic inflow modelling for rotorcraft flight
dynamics’. Vertica, 2, 3, 213–242.
Goldstein, S., (1929). ‘On the vortex theory of screw propeller’. Proc. Roy. Soc. (A), 123, 440.
HaQuang, N. and Peters, D. A., (1988). ‘Dynamic inflow for practical applications’. J. Am. Heli. Soc.,
Technical Note.
Hansen, M. O. L, Sorensen, N. N. and Flay, R. G. J. (2000) ‘Effect of placing a diffuser around a wind
turbine’, Wind Energy, vol 3, pp207–213
Himmelskamp, H., (1945). Profile investigations on a rotating airscrew. Ph.D. Thesis, Gottingen
University, Germany.
Hoerner, S. F., (1965). ‘Pressure drag on rotating bodies’. Fluid dynamic drag (ed. Hoerner, S. F.), pp.
3–14.
Hoerner, S.F. and Borst, H.V., 'Fluid Dynamic Lift', Published by Mrs. Hoerner, L.A., Hoerner Fluid
Dynamics, Vancouver, WA , USA, 1985.
Holm, R and Gustavsson, J., 'A PIV Study of Separated Flow around a 2D Airfoil at High Angles of
Attack in a Low Speed Wind Tunnel', in proceedings of IEA Annex XI, 'Aerodynamics of Wind
Turbines', 13th symposium, November 1999, Stockholm.
Holten, van Th., 'On the Validity of Lifting line Concepts in Rotor Analysis', Vertica, Vol. 1 pp. 239-
254, 1976.
Prandtl, L. and Tietjens, O. G. (1957) Applied Hydro and Aeromechanics, Dover Publications, New
York.
Schlichting, H. (1968) Boundary-Layer Theory, McGraw-Hill, New York.
Simpson, R.L., 'Aspects of Turbulent Boundary Layer Separation', Prog. Aerospace Sci., Vol. 32, pp.
457-521, 1996.
White, F. M. (1991) Viscous Fluid Flow, McGraw-Hill, New York.
Kuik, G.A.M. van, ‘On the Limitations of Froude’s Actuator Disk Concept’, ph.d. thesis 1991.

| Wind –Turbine Performance Pavesi - 150 -


9 Power Control
A wind turbine is essentially a device that captures part of the wind energy and converts it into
useful work. In particular, wind energy control system connected to electric power networks must be
designed to minimise the cost of supplied energy ensuring safe operation as well as acoustic emission
and power quality standards.
The minimisation of the energy cost involves a series of partial objectives.
These objectives are actually closely related and sometimes conflicting. Therefore, they should
not be pursued separately. Conversely, the question is to find a well balanced compromise among them.
These partial goals can be arranged in the following topics:
 Energy capture: Maximisation of energy capture taking account of safe operation restrictions
such as rated power, rated speed and cut-out wind speed , et c.
 Mechanical loads: Preventing the turbine from excessive dynamic mechanical loads. This
general goal encompasses transient loads alleviation, high frequency loads mitigation and
resonance avoidance.
 Power quality: Conditioning the generated power to comply with interconnection standards.

9.1.1 Energy Capture


For a wind turbine, the generation capacity specifies how much power call be extracted from the
wind taking into consideration both physical and economic constraints. It is usually represented as a
curve on the generated power – wind speed plane, the so-called ideal power curve.
The ideal power curve for a typical wind
turbine is sketched in Figure 8.4.1. The range of
operational wind speeds is delimited by the cut-
in (Vmin) and cut-out (Vmax) wind speeds. Below
cut-in wind speed, the available wind energy is
too low to compensate for the operation costs
and losses. Above cut-out wind speed, the
turbine is shut down to prevent from structural
overload. Constructing the turbine robust enough
to support the underlying mechanical stresses
under very high wind condition is completely Figure 8.4.1 Ideal Power Curve
uneconomical. In fact, even though wind speed
above Vmax contain huge energy, its contribution to the annual average energy is negligible
The ideal power curve remains constant at rated power PN above the rated wind speed VN.
The ideal power curve exhibits three different regions with distinctive generation objectives. At
low wind speeds (region I), the available power is lower than rated power. The available power is
defined as the power in the wind passing through the rotor area multiplied by the maximum power
coefficient CPmax.
The generation objective in region I is to extract all the available power. Therefore, the ideal
power curve in this region follows a cubic parabola defined by wind available power. On the other side,
the generation goal in the high wind speed region (region III) is to limit the generated power below its
rated value to avoid over loading. In this region the available power exceeds rated power, therefore the
turbine must be operated with efficiency lower than CPmax· Finally, there is region II which is actually a
transition between the optimum power curve of region I and the constant power line of region III. In

| Power Control Pavesi - 151 -


this region, rotor speed is limited to maintain acoustic noise emission within admissible level and to
keep centrifugal forces below values tolerated by the rotor.

9.1.2 Mechanical Loads


Keeping in mind the minimisation of the energy cost, the control system should not merely
designed to track as tightly as possible the ideal power curve. In fact, the other control objectives must
not be ignored. For instance, the mechanical loads wind turbines are exposed to must also be
considered. Mechanical loads may cause fatigue damage on several devices, thereby reducing the
useful life of the system. Since the overall cost of the wind turbine is therefore spread over a shorter
period of time, the cost of energy will rise.
There are basically two types of mechanical loads, namely static and dynamic ones. Static loads
result from the interaction of the turbine with the mean wind speed. Much more important from the
control viewpoint are the dynamic loads, which are induced by the spatial and temporal distribution of
the wind speed field over the area swept by the rotor. Dynamic loads comprise variations in the net
aerodynamic torque that propagate down the drive-train and variations in the aerodynamic loads that
impact on the mechanical structure. They are the so-called drive-train and structural loads, respectively.
There is also another common classification of dynamic loads. On the one hand there are the
transient loads, which are induced by turbulence and gusts. They are predominantly of low frequency.
Transient loads have very important implications in high wind speeds, particularly for the
determination of the components rating. The transition between maximum power tracking (region I)
and power regulation (region III) and the way power is limited in above rated wind speeds have a direct
impact on transient load. Therefore, the planning of the control strategy must also take them into
consideration.
In addition, controller setup and design also influence the transient loads. In fact, the tighter the
closed-loop system follows the steady-state control strategy curve after a wind gust, the heavier the
transient loads will be.
On the other hand, rotational sampling induces high-frequency cyclic loads concentrated around
spectral peaks at multiples of rotor speed. In this respect, control systems are increasingly important as
the wind turbines are larger and their components more flexible.
Cyclic loads are highly influenced by the control strategy as well as by the controller setup and
design. The control of the electric generator affects the propagation of drive-train loads whereas the
pitch control impacts directly on the structural loads.

9.1.3 Power Quality


Power quality affects the cost of energy in several ways. For instance, poor power quality may
demand additional investments in power lines, or may impose limits to the power supplied to the grid.
Because of the long-term and short-term variability of the energy resource and the interaction
with the power network, wind generation facilities are conventionally considered as poor quality
suppliers. Therefore, the control system design must also take power conditioning into account. This
control requirement is more and more relevant as the power scale of wind generation facilities
approaches the output rating of conventional power plants. Power quality is mainly assessed by the
stability of frequency and voltage at the point of connection to the grid and by the emission of flicker.
Commonly, when connected to the bulk network, single wind turbines or small-scale wind farms
do not affect the frequency. However, this is not the case when the wind turbine is part of an isolated
power system or when we are dealing with a large-scale wind farm.

| Power Control Pavesi - 152 -


On the one hand, slow voltage excursions take place when the power extracted by the wind
turbine changes with mean wind speed. With modern wind turbines being connected to grid through
power converters, the current tendency is to take advantage of the control flexibility provided by the
power electronics.
On the other hand, the cyclic loads originated by rotational sampling effects and propagated down
the drive-train towards the grid produce fast fluctuations of grid voltage. Regretfully, the frequency of
these cyclic loads may fall into the range of human eye sensitivity, thereby they induce flicker on the
electric lines. Flicker is defined as an impression of unsteadiness sensation induced by a fluctuating
light stimulus that causes consumer annoyance. These voltage fluctuations and flicker can be attenuated
by including passive or active filters, or, in the case of variable-speed wind turbine, by controlling the
reactive power handled by the electronic converters.

9.2 Modes of Operation


A control system reaches its steady-state operating point when the net torque applied to the
system is zero, i.e., when the generator reaction torque equals the aerodynamic torque developed on the
rotor (both torques referred to the same side of the gearbox). This steady-state condition is illustrated in
Figure 9.2.1. There, the aerodynamic torque characteristic of the rotor is plotted for different values of
wind speed and pitch angle. Also, the reaction torque characteristic of the generator is shown for two
values of the zero-torque speed
,. Clearly, the point where the
reaction torque intersects the
aerodynamic torque curve for a
given wind speed is the operating
point at that wind speed.
For instance, given 0 and
P P1 is the operating point at
wind speed Vl P2 is the operating
point at wind speed V2 and so on.
It turns out from Figure 9.2.1.
that either the generator or the
aerodynamic characteristic must
be modified to operate the turbine
at other operating points. This
sort of flexibility is usually
necessary to accomplish the
Figure 9.2.1 Operating points for different operating conditions
control objectives.
Figure 9.2.1 illustrates how the operating point shifts from P1 to Q1 when the zero-torque speed 
controlled by the converters increases from P to Q that is at variable speed. This mode of operation is
useful, for instance, to track the maximum power locus as wind speed varies below its rated value.
On the other hand, pitch control is the most popular way to change the aerodynamic torque
response of the turbine. Figure 8.2.4 plots how the aerodynamic torque characteristic varies with pitch
angle. Figure 9.2.1 shows also how the operating point is modified even though wind speed remains
constant. In fact, the operating point changes from P1 to R1 when pitch angle increases. Additionally, it
can be observed that the operating point can be maintained constant despite wind changes by means of
a. suitable pitch angle adjustment. In the figure, the operating point P2 for V2 and coincides with the

| Power Control Pavesi - 153 -


operating point R1 for Vl and R. Variable-pitch operation is particularly useful to allows the turbine to
keep operating at a fixed point despite wind speed fluctuations.
The term 'modes of operation' denote the various ways wind turbines can be programmed to
work. They are essentially determined by the feasible ways to actuate on the turbine. Fixed-speed,
variable-speed, fixed-pitch and variable-pitch are the commonest ones. Since wind turbines work under
different conditions, these modes of operation are usually combined to attain the control objectives
over the full range of operational wind speeds. Accordingly, wind turbines can be classified into four
categories, namely:
 Fixed-speed fixed-pitch (FS-FP).
 Fixed-speed variable-pitch (FS-VP).
 Variable-speed fi.'Ced-pitch (VS-FP) .
 Variable-speed variable-pitch (VS-VP) .
Obviously, the capability of a wind energy control system to satisfy the control objectives closely
depends on the flexibility of its operation modes.

9.3 Control Strategies


In some way, the control strategy describes how the turbine is programmed to approach in steady-state the
ideal power curve in the power - wind speed plane (Figure 8.4.1). Thus, the control strategy settles the steady-
state values of torque (or power) and rotor speed for each wind speed within the range of turbine operation. The
control strategy affects the controller setup and design.
Despite its stationary nature, the control strategy has a strong influence on the dynamic behaviour of the
wind electric system control. In particular, the planning of the transition between maximum power tracking and
power regulation has a direct impact on the transient loads. Also, operation in some region may excite an un-
damped vibration mode. Consequently, the selection of the desired operating points must arise from a
compromise between the control objectives. For instance, a transition region smoother than the ideal one
alleviates significantly the transient loads at the cost of some loss in energy capture.

9.3.1 Fixed-speed Fixed-pitch


Fixed-speed fixed-pitch operation has been the dominant configuration during some decades. However,
the number of commercial wind turbines based on this concept has declined lately.
In this scheme, the asynchronous electric machine is directly coupled to the power network. Thus,
its torque characteristic cannot be modified . Consequently, the generator speed is locked to the power
line frequency. By this reason , it is said that the wind energy system control operates at fixed speed. In
reality, the speed varies a few percent along the torque characteristic of the generator because of the
slip. Since no extra hardware is purposely added to implement the control strategy, FS-FP are very
simple and low-cost. As an adverse consequence, their performance is rather poor. In fact, no active
control action can be done to alleviate mechanical loads and improve power quality. Further,
conversion efficiency is far from optimal.
Figure 9.3.1 illustrates the basic control strategy for FS-FP wind turbines in the torque - rotational
speed plane. The solid line depicts the reaction torque characteristic, whereas the grey lines represent
the aerodynamic torque characteristics for different wind speeds between Vmin and Vmax.
The points of intersection are the steady-state operating points at the corresponding wind speeds.
Since neither the reaction torque nor the aerodynamic torque characteristics can be modified, all
feasible operating points are constrained to the segment FD. Since rotational speed is almost constant
along FD, power is more or less proportional to torque all along this operating locus. The operating

| Power Control Pavesi - 154 -


point F corresponds to the cut-in
wind speed Vmin whereas D is the
point where the reaction torque
characteristic intersects the stall
front, that is the upper boundary of
all aerodynamic torque
characteristics. This point
determines the maximum power
that can be extracted by the
turbine. At wind speed V0, for
which the aerodynamic torque
characteristic passes through point
D, the turbine stalls. Therefore,
higher wind speeds lead to lower
aerodynamic power.
That is why the operating
point moves back along the
generator torque characteristic Figure 9.3.1 Fixed-Speed Fixed Pitch control strategy.
until point G related to the cut-out
wind speed Vmax is reached. It is worthy to remark that there exists a superposition of operating points
in the segment GD. This means that the wind speed cannot be uniquely determined from the operating
point.
The parabola plotted in Error! Reference source not found. depicts the locus of maximum
conversion efficiency, also called maximum power or CPmax locus. It shows up that the turbine operates
with maximum efficiency at a unique wind speed VE. This situation corresponds to point E where the
maximum power locus, the reaction torque and the aerodynamic torque for VE intersect. At this is
point, wind and rotational speeds satisfy 0 = R/VE.
Suppose now that the wind turbine is operating at E when wind speed drops from VE to VJ. Then,
the new operating point is J , being J  E. At this point the tip-speed-ratio . j = R/Vj is higher than
0, therefore the conversion efficiency decreases. In order to capture all the power available at wind
speed VJ, the turbine should be operated at point J' where the rotational speed J' = 0 VJ/ R is lower.
This is impracticable with a fixed-speed control strategy.
FS-FP wind turbines are stall regulated at high wind speeds. That is, power limitation below rated
power is accomplished by passive stall. Therefore, the selection of the control strategy comes down to
choose the transmission ratio of the gearbox for the generator characteristic to pass through point D.
Note that this point is the intersection between the rated power hyperbola (which is the boundary of the
safe operating region) and the stall front. So, rated power is exceeded at no wind speed.
Figure 9.3.2 helps to understand the passive stall method for power limitation. The figure
qualitatively shows the forces acting on a blade element before (grey) and after (black) stall. Recall that
rotational speed and pitch angle are fixed . Then, the angle of attach  increases as the wind speed
experienced by the blade element rises from V0 to V1 . When  exceeds a given value, the airflow
separates from the upper side of the aerofoil.
This reduces the lift and increases abruptly the drag. Changes in lift fL and drag fD forces lead to
a large increment of the axial thrust force fT whereas the tangential force fr decreases slightly. As a
result, the aerodynamic torque and power decrease. It is said that the turbine stalls. An undesirable

| Power Control Pavesi - 155 -


consequence of stall regulation is the
increased thrust that causes heavier
aerodynamic loads.
Figure 9.3.3 illustrates the power
capture features of the basic FS-FP control
strategy shown in Figure 9.3.1. The top
part of the figure compares the actual and
ideal power curves, whereas the bottom
part shows the conversion efficiency
versus. wind speed. The points marked in
the figure are in correspondence with those
of Figure 9.3.1. It can be observed that the
extracted power does not match the ideal
power curve. This means lower energy
capture. In the low wind speed region, the
turbine operates with maximum efficiency
only at one point (point E). In above rated
wind speeds, power regulation is rather
poor. It is seen that rated power is attained
only at one wind speed (point D) whereas
power decreases at lower and higher wind
speeds. This poor regulation is put down to
the lack of flexibility of the operation
mode.
In addition to the low conversion
efficiency, fixed-speed operation suffers
from other shortcomings more related with
the dynamic behaviour· of the wind energy
control system. For instance, the poor
regulation property translates into active Figure 9.3.2 Passive stall strategy for power limitation:
and reactive power fluctuations on the (a) forces on a blade element and (b) drag and lift
power lines. Additionally, this mode of coefficients
operation does not provide control actions
to inject damping to the drive-train, which would attenuate high-frequency loads and flicker emission.
Actually, the decline of FS-FP wind turbines is mainly due to their bad power quality, rather than to
reduced energy capture.

| Power Control Pavesi - 156 -


Figure 9.3.3 Basic fixed-speed fixed-pitch control strategy: (a)
output power and (b) conversion efficiency vs. wind speed

9.3.2 Fixed-speed Variable-pitch


Active pitch control achieves power limitation above rated wind speed by rotating all or part of
each blade about its axis in the direction which reduces the angle of attack and hence the lift coefficient
a process known as blade feathering. The main benefits of active pitch control are increased energy
capture, the aerodynamic braking facility it provides and the reduced extreme loads on the turbine
when shut-down.
The pitch change system has to act rapidly, i.e., to give pitch change rates of 5°/s or better in
order to limit power excursions due to gusts enveloping the whole rotor to an acceptable value.
However, it is not normally found practicable to smooth the cyclic power fluctuations at blade passing
frequency due to blades successively slicing through a localized gust with the result that the large
power swings of up to about 100 percent can sometimes occur.
The extra energy obtainable with pitch control is not all that large. A pitch regulated machine
with the same power rating as a stall-regulated machine, utilizing the same blades and rotating at the
same speed will operate at a larger pitch angle below rated wind speed than the stall-regulated machine,
in order to reduce the angle of attack and hence increase the power output at wind speeds approaching
rated. If the 500 kW, 40 m diameter, 30 r.p.m. stall-regulated machine described in Section 6.5.3 is
taken as baseline, a 500 kW, 30 r.p.m. pitch-regulated machine utilizing the same blades at optimum
pitch would produce about 2 percent more energy. The optimum rotational speed is found to be about
33 r.p.m., which increases the energy gain to about 4 percent. The power curve of the 500 kW, 40 m
diameter pitch controlled machine rotating at 33 r.p.m. is compared with the power curve of the
corresponding stall-regulated machine utilizing the same blades, but rotating at 30 r.p.m. in Error!

| Power Control Pavesi - 157 -


Figure 9.3.4 Power Curves for Different Pitch Angles: 40 m Diameter Rotor Rotating at 33
r.p.m.
Reference source not found.. Note that the knee in the power curve at rated speed will be more
rounded in practice because the pitch control will not keep pace with the higher frequency components
of turbulence.
Figure 9.3.4 shows a family of power curves for a range of positive pitch angles for the 500 kW,
40 m diameter pitch-controlled machine rotating at 33 r.p.m.. The intersections of these curves with the
500 kW abscissa define the relationship between steady wind speed and pitch angle required for power
control. It is readily apparent from the power curve gradients at the intersection points that rapid
changes of wind speed will result in large power swings when the mean wind speed is high.
The range of blade pitch angles required for power control is typically from 0° (often referred to
as ‘fine pitch’), at which the tip chord is in the plane of rotation or very close to it, and about 35°.
However, for effective aerodynamic braking, the blades have to be pitched to 90° or full feather, when
the tip chord is parallel to the rotor shaft with the leading edge into the wind.
A variety of pitch actuation systems have been adopted.
They are divided between those in which each blade has its own actuator and those in which a
single actuator pitches all the blades. The former arrangement has the advantage that it provides two or
three independent aerodynamic braking systems to control overspeed, and the disadvantage that it
requires very precise control of pitch on each blade in order to avoid unacceptable pitch angle
differences during normal operation.
An advantage of the latter arrangement is that the pitch actuator, e.g. a hydraulic cylinder, can be
located in the nacelle, producing fore-aft motion of the pitch linkages in the hub by means of a rod
passing down the middle of a hollow low-speed shaft (see Figure 9.3.4). Alternatively, the axial

| Power Control Pavesi - 158 -


position of the rod can be controlled by means of a ball-screw and ball-nut arrangement, in which the
ball-nut is driven by a servomotor. Normally the ball-nut is driven at the same speed as the rotor, but
when a change of pitch is required the ball-nut rotational speed is altered temporarily. This system is
arranged to be fail-safe, so that should the servomotor or its control system fail, the servomotor is
braked automatically and the ball-nut drives the blade pitch to feather.
Where hydraulic cylinders are used to pitch blades individually, they are mounted within the hub
and each piston rod is usually connected directly to an attachment on the blade bearing (see Figure
9.3.5). The attachment point follows a circular path as the blade pitches, so the cylinder has to be
allowed to pivot. The alternative solution of employing an electric motor to drive a pinion engaging
with teeth on the inside of the blade bearing consequently appears rather neater (see Figure 9.3.7). Both
systems require a hollow shaft to accommodate either hydraulic hoses or power cables for pitch
actuation together with signal cables for pitch angle sensing. In addition, appropriate slip rings are
required at the rear end of the shaft.
Methods of providing back-up power supplies to ensure blade feathering in the event of grid-loss
are considered in next Section.
Although full-span pitch control is the option favoured by the overwhelming majority of
manufacturers, power control can still be fully effective even if only the outer 15 percent of the blade is
pitched. The principal benefits are that the duty of the pitch actuators is significantly reduced, and that
the inboard portion of the blade remains in stall, significantly reducing the blade load fluctuations. On

Figure 9.3.5 Pitch Linkage System Used in Conjunction with a Single Hydraulic
Actuator Located in the Nacelle. (The central triangular ‘spider’ is connected to the
actuator by a rod passing through the hollow low-speed shaft. Links from the spider
drive the blade pitch via braced arms cantilevering into the hub from each blade.
Each arm is parallel to its blade axis, but eccentric to it).

| Power Control Pavesi - 159 -


Figure 9.3.6 Blade Pitching System Using Separate Hydraulic Actuators for Each Blade.
(Each actuator cylinder is supported on gimbal-type mountings bolted to the hub, and its
piston applies a pitching torque to the blade via a cantilevered conical tube eccentric to the
blade axis. The blade is attached to the outer ring of the pitch bearing).
the other hand partial-span pitch control has several disadvantages as follows:
• the introduction of extra weight near the tip,
• the difficulty of physically accommodating the actuator within the blade profile,
• the high bending moments to be carried by the tip-blade shaft,
• the need to design the equipment for the high centrifugal loadings found at large radii,
• the difficulty of access for maintenance.
It should be apparent from the above brief survey of pitch actuation systems that the design of the
hardware required for pitch-regulation is a significant task.
Moreover, as regards the controller, pitch-regulation introduces the need for fast response closed
loop control, which is not required for the supervisory functions on a stall-regulated machine. Thus the
benefits of pitch control have to be weighed carefully against all the additional costs involved,
including the cost of maintenance.
Another factor that needs to be considered is fatigue loading. This increases significantly on full-
span pitch-regulated machines, because the rate of change of lift coefficient with angle of attack
remains at about 2π instead of reducing to zero as the blade goes into stall, with the result that rapid
changes in wind speed above rated will cause bigger thrust load changes.
The FS-VP mode of operation has also often been employed in commercial wind turbines during
past decades, particularly in medium to high power. Fixed-speed operation means that maximum power
conversion is attainable only at a single wind speed. Therefore, conversion efficiency below rated wind

| Power Control Pavesi - 160 -


Figure 9.3.7 Blade Pitching System Using a Separate Electric Motor for Each Blade. (A
pinion, driven by the motor via a planetary gearbox, engages with gear teeth on the inside of
the inner ring of the pitch bearing, to which the blade is bolted. The blade is not attached to
the bearing in this photograph, so the fixing holes are visible).
speed cannot be optimised. This kind of turbine is usually programmed to operate at fixed pitch below
rated wind speed . However , variable-pitch operation in low wind speeds could be potentially helpful
to enhance somewhat the energy capture. In above rated wind speeds, power is limited by continuously
adjusting the pitch angle. There are basically two methods of power regulation by pitch control, namely
pitch-to-feather and pitch-to-stall. The former method is conventionally referred to as pitch angle
control , whereas the second method is also known as active stall or combi stall. To avoid
misunderstandings, we will refer to them as pitch-to-feather and pitch- to-stall.

| Power Control Pavesi - 161 -


9.3.2.1 Power Limitation by Pitch-to-feather

This method essentially consists in


feathering the blades as wind speed rises.
Thus, it is based on an aerodynamic
phenomenon completely different to stall.
Figure 9.3.8 illustrates the method. In the top
part of the figure , the forces acting on a blade
element that experiences axial wind speeds V0
and V1>V0 are depicted by grey and black
vectors, respectively. When wind speed rises
from V0 to V1, the angle  that the relative
airflow makes with the rotor plane increases.
As response, the controller raises the pitch
angle  so that the angle of attach  decreases.
As a result, the lift coefficient CL drops
whereas the drag coefficient CD remains low.
It can be thought that the controller adjusts the
lift force fL in order to keep the force fr in the
rotor plane constant. It can be observed in the
figure that, contrary to what happens with
passive stall, the thrust force fr decreases as
wind speed rises.
Since thrust efforts produce
aerodynamic loading on the structure, this is a
significant advantage of this method. As a
disadvantage, pitch-to-feather regulation
requires a considerable control effort since
large pitch changes are necessary to
compensate for wind power fluctuations. Figure 9.3.8 Pitch-to-feather strategy for power
limitation: (a) forces on a blade element and (b) drag
9.3.2.2 Power Limitation by Pitch-to-stall and lift coefficients

The alternative to pitch-to-feather is pitch-to-stall. In this case, the pitch angle is adjusted in the
opposite direction. In fact, the pitch angle is reduced in order to increase the incidence angle, rather
than to decrease it, as wind speed rises. That is, the pitch angle is controlled to actively induce stall
above rated wind speed. Thanks to the control flexibility, this method holds better regulation features
than passive stall. Figure 9.3.9 illustrates the method. The top part of the figure depicts the forces
acting on a blade element that experiences axial wind speeds V0 (grey) and V1 (black), with V1> V0 .
The angle  increases with wind speed. As a result, the angle of attach  tends to increase (recall
Figure 9.3.2 explaining passive stall). In pitch-to-stall, the pitch angle  is reduced to increase further
the incidence angle, hence reinforcing stall.
Lift drops whereas drag rises abruptly. The composition of these forces results in a force in the
rotor plane that remains constant, thus maintaining the aerodynamic power at its rated value. A
disadvantage of this method , as well as of all methods based on stall, is that the thrust force increases

| Power Control Pavesi - 162 -


drastically as the turbine goes into stall.
This translates into heavy aerodynamic
loads.
As main attractive feature, this
method requires a comparatively low
control effort to regulate power. It is
worthy to note that it is no more necessary
for the control strategy to pass exactly
through the point of intersection between
the rated power hyperbola and the stall
front. This gives in principle more
freedom to design the transmission ratio.
Active stall control achieves power
limitation above rated wind speed by
pitching the blades initially into stall, i.e.,
in the opposite direction to that employed
for active pitch control, and is thus
sometimes known as negative pitch
control. At higher wind speeds, however,
it is usually necessary to pitch the blades
back towards feather in order to maintain
power output at rated.
A significant advantage of active
stall control is that the blade remains
essentially stalled above the rated wind
speed, so that gust slicing results in much
smaller cyclic fluctuations in blade loads
and power output. It is found that only
small changes of pitch angle are required
to maintain the power output at rated, so Figure 9.3.9 Pitch-to-stall strategy for power limitation:
pitch rates do not need to be as large as for (a) forces on a blade element and (b) drag and lift
positive pitch control. coefficients
Moreover, full aerodynamic braking requires pitch angles of only about -20°, so the travel of the
pitch mechanism is very much reduced compared with positive pitch control.
Figure 9.3.10 compares schedules of pitch angle against wind speed for active stall control and
active pitch control for the same blade. The active stall control schedule is derived from the intersection
of the family of power curves for different negative pitch angles with the 500 kW abscissa in Figure
9.3.11, while the active pitch control schedule is derived from Figure 9.3.4
The principal disadvantage of active stall control is the difficulty in predicting aerodynamic
behaviour accurately in stalled flow conditions.

| Power Control Pavesi - 163 -


Figure 9.3.10 Specimen Pitch Angle Schedules for Active Pitch Control and Active Stall
Control

Figure 9.3.11 Power Curves for Different Negative Pitch Angles: 40 m Diameter Rotor
Rotating at 33 r.p.m.

| Power Control Pavesi - 164 -


Figure 9.3.12 Basic fixed-speed variable-pitch (pitch-to- feather)
control strategy

9.3.2.3 Basic Control Strategy

Figure 9.3.9 depicts the basic


pitch-to-feather control strategy
for FS-VP wind turbines. The
aerodynamic torque characteristic
for different wind speeds within
the operational envelope is also
plotted with pitch angle as a
parameter. Since there is no
control on the reaction torque,
rotational speed is almost constant
all along the operating locus.
Below rated wind speed, the
control strategy is similar to the
FS-FP one treated in 9.3.1. That is,
the turbine is operated at some
point in the segment FD. Above
rated wind speed, the controller
adjusts the pitch angle to regulate
aerodynamic power at rated, that
is, the turbine is programmed to
operate at point D. It is observed
how the aerodynamic torque Figure 9.3.13 Basic fixed-speed pitch-to-feather and pitch-to-stall
characteristics are modified to control strategies: (a) output power and (b) power efficiency vs.
wind speed

| Power Control Pavesi - 165 -


pass through this point D.
In the case of pitch-to-stall, all trajectories pass through point D also, but the aerodynamic torque
characteristics are shaped differently. This case can be interpreted as that the stall front moves
rightwards as wind speed increases.
Figure 9.3.12 shows the power regulation features of the basic control strategies for FS-VP wind
turbines. The points marked in the figure are in correspondence with those of Figure 9.3.9. Since the
operating points of the wind turbine on the torque - speed plane are the same no matter the control
strategy is pitch-to-feather or pitch-to-stall, Error! Reference source not found. is valid for both
cases. The top part of the figure compares the actual and ideal power curves whereas the bottom part
shows the conversion efficiency vs. wind speed . Below rated wind speed , the real power curve is
similar to the one for the FS-FP control strategy plotted in Figure 9.3.3.
Even though maximum conversion efficiency is attained only at wind speed VE, variable-pitch
wind turbines are designed to have near optimum conversion efficiency over a wide range of wind
speeds below VN. Above rated wind speed , ideal power regulation is achieved provided there are not
pitch actuator constraints. That is, the theoretical power curve is attached to the ideal curve (compare
with Figure 9.3.3). The bottom part of the figure illustrates how the power coefficient is shaped above
rated wind speed to match the ideal power curve.

9.3.3 Variable-speed Fixed-pitch


The variable-speed alternative to fixed-speed has become popular in commercial wind turbines,
particularly for operation in low wind speeds. The benefits commonly ascribed to variable-speed
operation are larger energy capture, dynamic loads alleviation and power quality enhancement. With
wind energy currently reaching large penetration factors, the demands for power quality improvement
gave a decisive impetus to the use of variable-speed schemes. Maximum efficiency conversion is
achieved at ( = 0 and  = 0. Therefore, to maximise energy capture below rated power, both the
pitch angle and the tip-speed-ratio must be kept constant at these values.
In particular, the condition  = 0 means that the rotor speed must change proportionally to wind speed,
i.e .,
V
r0  0 (10.3.1)
r
As it is known, variable-speed operation requires decoupling the rotational speed from the line
frequency. A suitable control of the underlying electronic converters produces parallel displacements of
the generator torque characteristic towards higher or lower speeds. Thus, the turbine can be controlled
to operate at different points, for instance to track the optimum speed as wind speed fluctuates. The
maximum power locus is the equation of a parabola in the torque versus speed plane.
In low wind speeds, variable-speed wind turbines are controlled to track more or less tightly this
CPmax locus. Thus, variable-speed control strategies essentially differ in the way power is limited above
rated wind speed. For VS-FP wind turbines, there are basically two approaches based on passive and
speed-assisted stall, respectively. In this subsection we restrict our analysis to these basic VS-FP
control strategies.

| Power Control Pavesi - 166 -


9.3.3.1 Variable Speed with Passive Stall Regulation

This control strategy is


identified by the points AEDG in
Figure 9.3.14. Clearly, the turbine
works at two different modes
throughout its operating region. In
the low wind speed region , more
precisely between Vmin and VE,
the turbine is programmed to
operate along the quadratic curve
AE. That is, the c:ont.rol strategy
coincides with the CPmax locus.
Clearly, the turbine is operated at
variable speed in this region. For
wind speeds above VE the
operating point of the turbine
moves along the segment ED.
That is, the turbine is operated at
fixed speed in this wind speed
Figure 9.3.14 Basic variable-speed fixed-pitch control strategies
region, hence performing like an
with passive (AEDG) and speed-assisted (ABCDG') stall regulation
FS-FP one. Therefore, power is
limited by passive stall as
described in Figure 9.3.2.
Again, there exists a
superposition of operating
points before and after stall in
the segment GD. Note that
the generator torque
characteristic containing the
segment ED passes through
the intersection point between
the rated power hyperbola
and the stall front. This is to
capture as much energy as
possible without over
loading. Figure 9.3.15 shows
the power capture properties
of this control strategy. The
top part of the figure
compares the actual (black)
and ideal (grey) power
curves, whereas the bottom
part displays the conversion Figure 9.3.15 Variable-speed fixed-pitch control strategies with passive
efficiency (black) versus the stall regulation (AEDG) and speed-assisted stall regulation (ABCDG'):
(a) captured power and (b) power efficiency vs. wind speed

| Power Control Pavesi - 167 -


wind speed. The points A, E , D and G are in correspondence with those of Figure 9.3.1. It is observed
that the real power curve is attached to the ideal one between Vmin and VE . Thus, power conversion is
maximised in this speed range. At point E, the real power curve separates from the ideal one, leading to
some loss in captured power. The poor regulation features of passive stall and other drawbacks of
fixed-speed operation have already been mentioned.

9.3.3.2 Variable Speed with Assisted Stall Regulation

This control strategy can be identified by the points ABCDG' in Figure 9.3.14. The wind turbine
is operated at variable speed throughout its operational range. The control strategy coincides with the
CPmax locus from A to B. All along this curve, rotational speed increases proportionally to wind speed
until rated speed DN is reached at B. Since N cannot be exceeded , the operating point moves along
the segment BC as wind speed increases from VN to VN. That is, the operating speed of the WECS
remains constant in this wind speed region . For wind speeds above rated, the operating point moves
along the rated power hyperbola towards the stall front . At wind speed VD, the control strategy reaches
the stall front, i.e., the turbine stalls. For higher wind speeds the operating point moves back along the
hyperbola until shut-clown at G’ It turns out that this control strategy accomplishes the ideal power
curve of Figure 8.4.1. This is corroborated in Figure 9.3.15 where the distinctive points of the control
strategy are marked on the ideal power curve. The three regions are clearly distinguished in the torque -
rotational speed graph. Although this control strategy potentially enables maximum energy capture, it
also brings with it some transient problems . Problems arise when the WECS operates under turbulent
conditions around its rated operating point (point C). For instance, suppose the turbine is being
operated at C and wind speed increases suddenly above VN. Aerodynamic power tends to increase in
consequence. To compensate for this increment, the operating point is moved leftwards along the rated-
power hyperbola, thus leading to a reduction of rotational speed. Accordingly, part of the kinetic
energy stored in the rotor is reduced. Necessarily, the energy in excess is transmitted thought the drive-
train and supplied to the AC grid. Thus, operation around the nominal point inevitably leads to
undesirable transient loads and electric power fluctuations that degrade power quality. The amplitude
of aerodynamic power fluctuations and, hence, of transient loads decreases as the operating point
approaches the stall front.

9.3.4 Variable-speed Variable-pitch


Variable-speed variable-pitch control strategies are being more and more common in commercial
wind turbines. In this scheme, the turbine is programmed to operate at variable speed and fixed pitch
below rated wind speed and at variable pitch above rated wind speed. Both pitch-to-feather and pitch-
to-stall strategies can be followed. Figure 9.3.16 depicts the basic variable speed pitch-to-feather
control strategy on the torque- rotational speed plane. In low wind speeds, the turbine is operated along
the CPmax locus between the points A and B. At point B, the rotational speed gets to its upper limit N.
Therefore, rotational speed is regulated at this value on the segmeut BC as wind speed increases
from VN to VN. Above rated wind speed, the pitch angle is controlled in order to keep the turbine
operating at point C. Note that the segment BC comes down to point C' when the intersection between
the rated power hyperbola and the maximum efficiency parabola lies to the left of the rotational speed
limit. In this case, the operating locus is reduced to the curve AC'.
Variable-speed operation increases the energy capture at low wind speeds whereas variable-pitch
operation enables an efficient power regulation at higher than rated wind speeds. Note that this control

| Power Control Pavesi - 168 -


strategy also achieves the
ideal power curve of Figure
8.4.1. In addition, variable-
pitch operation alleviates
transient loads. This is an
important advantage of this
control strategy in
comparison with VS-FP
ones, particularly for large-
scale wind turbines.
Moreover ,
simultaneous control of
pitch and speed above rated
wind speed provides
important benefits to the
dynamic performance of the
WECS under high wind
conditions. This will be
exploited in the last chapter
of the book. Figure 9.3.16 Basic variable-speed variable-pitch (pitch-to-feather)
control strategy

9.4 Yaw control


As most horizontal-axis wind turbines employ a yaw drive mechanism to keep the turbine headed
into the wind, the use of the same mechanism to yaw the turbine out of wind to limit power output is
obviously an attractive one. However, there are two factors which militate against the rapid response of
such a system to limit power: first, the large moment of inertia of the nacelle and rotor about the yaw
axis, and second, the cosine relationship between the component of wind speed perpendicular to the
rotor disc and the yaw angle. The latter factor means that, at small initial yaw angles, yaw changes of,
say, 10° only bring about reductions in power of a few percent, whereas blade pitch changes of this
magnitude can easily halve the power output. Thus active yaw control is only practicable for variable
speed machines where the extra energy of a wind gust can be stored as rotor kinetic energy until the
yaw drive has made the necessary yaw correction. This design philosophy has been exploited
successfully in Italy on the 60 m diameter Gamma 60 prototype, which has an impressive maximum
yaw rate of 88/s (Coiante et al., 1989).

9.5 Braking Systems


9.5.1 Independent braking systems—requirements of standards
DS 472 and the GL rules both require that a wind turbine shall have two independent braking
systems. On the other hand, IEC 61400-1 does not explicitly require the provision of two braking
systems (stating that the protection system shall include one or more systems capable of bringing the
rotor to rest or to an idling state), but it does require the protection system to remain effective even after
the failure of any non-safe-life protection system component.
IEC 61400-1 and the GL rules require that at least one of the braking systems should act on the

| Power Control Pavesi - 169 -


rotor or low-speed shaft, but DS 472 goes further and requires that one should be an aerodynamic
brake.
Normal practice is to provide both aerodynamic and mechanical braking. However, if
independent aerodynamic braking systems are provided on each blade, and each has the capacity to
decelerate the rotor after the worst-case grid loss, then the mechanical brake will not necessarily be
designed to do this. The function of the mechanical brake in this case is solely to bring the rotor to rest,
i.e., to park it, as aerodynamic braking is unable do this.

9.5.2 Active pitch control


Blade pitching to feather (i.e., to align the blade chord with the wind direction) provides a highly
effective means of aerodynamic braking. Blade pitch rates of 10° per second are generally found
adequate, and this is of the same order as the pitch rate required for power control. The utilization of
the blade pitch system for startup and power control means that it is regularly exercised with the result
that the existence of a dormant fault is highly unlikely.
In machines relying solely on blade pitching for emergency braking, independent actuation of
each blade is required, together with fail-safe operation should power or hydraulic supplies passing
through a hollow low-speed shaft from the nacelle be interrupted. In the case of hydraulic actuators, oil
at pressure is commonly stored in accumulators in the hub for this purpose.

9.5.3 Pitching blade tips


Blade tips which pitch to feather have become the standard form of aerodynamic braking for stall-
regulated turbines. Typically the tip blade is mounted on a tip shaft, as illustrated in Error! Reference
source not found., and held in against centrifugal force during normal operation by a hydraulic
cylinder. On release of the hydraulic pressure (which is triggered by the control system, or directly by
an overspeed sensor), the tip blade flies outwards under the action of centrifugal force, pitching to
feather simultaneously on the shaft screw. The length of the tip blade is commonly some 15 percent of
the tip radius.

| Power Control Pavesi - 170 -


The ability of the control system to trigger blade tip activation is of crucial importance. On a
number of early machine designs, the blade tips were centrifugally activated only, so there could be
long periods without overspeed events when they did not operate. As a result there was a risk of seizure
when operation was eventually required. With the now commonplace arrangement enabling the control
system to activate the tip as well, the system can be routinely tested automatically.
The penalty is that the low-speed shaft needs to be hollow to accommodate the feed to the
hydraulic cylinder.

9.5.4 Spoilers
Spoilers are hinged flaps, which conform to the aerofoil profile when retracted, and stick out at
right angles to it when deployed. However, although such devices have been used in the past, they have
to be of considerable length in order to decelerate the rotor adequately (Jamieson and Agius, 1990).
Moreover, unless the design allows for their operation to be regularly tested, there is a risk that they
will fail to deploy when actually needed.

9.5.5 Other devices


Various other devices have been suggested, such as
• ailerons,
• the sliding leading edge device, or SLEDGE, in which a length of leading edge at the tip slides
radially outwards,
• the flying leading edge device, or FLEDGE, in which the whole leading edge together with an
adjacent section of the camber face is pitched towards feather.
Jamieson and Agius (1990) and Armstrong and Hancock (1991) give useful surveys of these and
other aerodynamic braking devices, and note that the SLEDGE device, which utilizes only 2 percent or
3 percent of the blade area, is highly effective aerodynamically. Derrick (1992) examines the
capabilities of the SLEDGE and FLEDGE devices for both braking and power control in more detail.

Figure 9.5.1 Passive Control of Tip Blade, Using Screw on Tip Shaft and Spring

| Power Control Pavesi - 171 -


Despite their promise, these devices have not yet found commercial application.

9.6 Mechanical brake options


The duty of the mechanical brake need only be that of a parking brake on machines where the
aerodynamic brakes can be actuated independently.
However, on pitch-regulated machines where blade position is controlled by a single actuator, full
independent braking capability has to be provided by the mechanical brake. It is worth noting that
several manufacturers of stall regulated machines fitted with independent tip brakes ensure that the
mechanical brake can stop the rotor unassisted. This may be to satisfy requirements in certain countries
that two independent braking systems of a different type are provided.
A wind turbine brake typically consists of a steel brake disc acted on by one or more brake
callipers. The disc can be mounted on either the rotor shaft (known as the low-speed shaft) or on the
shaft between the gearbox and the generator (known as the high-speed shaft). The latter option is much
the more common because the braking torque is reduced in inverse proportion to the shaft speeds, but it
carries with it the significant disadvantage that the braking torques are experienced by the gear train.
This can increase the gearbox torque rating required by as much as ca 50 percent, depending on the
frequency of brake application. Another consideration is that the material quality of brake discs
mounted on the high-speed shaft is more critical, because of the magnitude of the centrifugal stresses
developed.
The brake callipers are almost always arranged so that the brakes are spring applied and
hydraulically retracted, i.e., fail-safe.
Aerodynamic braking is much more benign than mechanical braking as far as loading of the blade
structure and drive train is concerned, so it is always used in preference for normal shut-downs.

9.6.1 Parking versus idling


Although a mechanical parking brake is essential for bringing the rotor to rest for maintenance
purposes, many manufacturers allow their machines to idle in low winds and some do so during high
wind shut-downs. The idling strategy has two clear advantages: it reduces the frequency of imposition
of braking loads on the gear train, and gives the impression to members of the public that the turbine is
operating even when it is not generating. On the other hand, gearbox and bearing lubrication must be
maintained throughout.

9.7 Fixed Speed, Two-speed or Variable-speed


Wind turbine rotors develop their peak efficiency at one particular tip speed ratio, so fixed speed
machines operate sub-optimally, except at the wind speed corresponding to this tip speed ratio. Energy
capture can clearly be increased by varying the rotational speed in proportion to the wind speed so that
the turbine is always running at optimum tip speed ratio, or alternatively a slightly reduced
improvement can be obtained by running the turbine at one of two fixed speeds so that the tip speed
ratio is closer to the optimum than with a single fixed speed.
Noise considerations are often of greater significance than energy capture in the decision to opt
for non-fixed speed operation. The aerodynamic noise generated by a wind turbine is approximately
proportional to the fifth power of the tip speed. Both variable speed and two-speed operation allow the
rotational speed to be substantially reduced in low winds, thus reducing turbine aerodynamic noise
dramatically when it could otherwise be objectionable because of low ambient noise.

| Power Control Pavesi - 172 -


9.7.1 Two-speed operation
At one time, two-speed operation was relatively expensive to implement because separate
generators were required for each speed of turbine rotation. Either generators of differing numbers of
poles were connected to gearbox output shafts rotating at the same speed, or generators with the same
number of poles were connected to output shafts rotating at differing speeds. The rating of the
generator for low-speed operation would normally be much less than the turbine rating.
The development of induction generators with two sets of independent windings has allowed the
number of poles within a single generator to be varied by connecting them together in different ways.
Standard generators of this type are now available which can be switched between four- and six-pole
operation, giving a speed ratio of 1.5. Given correct selection of the two operating speeds, this ratio
produces close to the optimum increase of energy capture.
It is important to note that, in the case of stall-regulated machines, only limited energy gains are
to be had by converting from fixed-speed to two-speed operation for a given rated power, because the
maximum rotational speed of the two-speed machine is restricted to the rotational speed of the fixed
speed machine in order to limit the power to the required value. Energy gains are only of the order of 2
or 3 percent, but, nevertheless, two-speed operation is often deemed worthwhile on stall regulated
machines because of noise considerations.
In the case of pitch-regulated machines, the energy gain obtainable by moving to two-speed
operation depends on the power rating and rotational speed of the baseline fixed speed machine. Where
these parameters have been chosen to be close to the optimum, in relation to rotor diameter and rotor
chord respectively, an energy gain of only about 3 percent is attainable. However, energy gains of up to
10 percent are possible when the baseline design is sub-optimal.
Some disadvantages of two-speed operation are:
 additional generator cost,
 extra switchgear is required, which is subjected to a demanding duty in terms of frequency
of operation,
 control of turbine speed is required at each speed change,
 energy is lost while the generator is disconnected during each speed change.

9.7.2 Variable-speed operation


By interposing a frequency converter between the generator and the network, it is possible to
decouple the rotational speed from the network frequency. As well as allowing the rotor speed to vary,
this also allows the generator air-gap torque to be controlled.
Variable-speed operation has a number of advantages:
 below rated wind speed, the rotor speed can be made to vary with wind speed to maintain
peak aerodynamic efficiency;
 the reduced rotor speed in low winds results in a significant reduction in aerodynamically-
generated acoustic noise – noise is especially important in low winds, where ambient wind
noise is less effective at masking the turbine noise;
 the rotor can act as a flywheel, smoothing out aerodynamic torque fluctuations before they
enter the drive train – this is particularly important at the blade passing frequency;
 direct control of the air-gap torque allows gearbox torque variations above the mean rated
level to be kept very small;
 both active and reactive power can be controlled, so that unity power factor can be

| Power Control Pavesi - 173 -


maintained – it is even possible to use a variable speed wind farm as a source of reactive
power to compensate for the poor power factor of other consumers on the network;
variable speed turbines will also produce a much lower level of electrical flicker.
In practice, losses in the frequency converter may amount to several per cent, counteracting the
increased aerodynamic efficiency below rated. In terms of energy capture, there is often little to choose
between a two-speed and a variable-speed machine. The load reduction possibilities, however, mean
that most large MW-scale turbines now use variable speed in some form. Variations in aerodynamic
torque at blade passing frequency are particularly significant in larger turbines because of the size of
the rotor compared to the lateral and vertical length scales of turbulence.
Clearly there is a significant cost associated with the variable-speed drive or frequency converter,
which must be weighed
against the advantages. Other
drawbacks include increased
complexity, although there is
no particular evidence of
reduced availability due to
power converter problems,
and the generation of electrical
noise and harmonics by the
inverter system. Modern
PWM inverters produce much
lower levels of undesirable
harmonics than earlier devices
because of the high switching
frequencies which can be
achieved. Electrical noise can
be a problem for control
signals within the turbine if Figure 9.7.1 Rotor efficiency and annual energy production versus
insufficient care is taken to rotor tip-speed ratio.
shield cables. Fibre optic transmission is increasingly being used, and this is not affected.
There are two principal methods of achieving variable speed operation: ‘broad range’ variable
speed, in which the generator stator is connected to the network via the frequency converter, and
‘narrow range’ variable speed where both the generator stator and rotor are connected to the network,
the stator directly, and the rotor through slip rings and a frequency converter.
Broad range variable speed allows the speed to vary under load from zero to the full rated speed,
but all the power output has to pass through the frequency converter. Narrow range variable speed
allows a much cheaper frequency converter, since only a fraction of the power passes through it, but
the speed can only vary by a more limited amount, typically 30–50% either side of synchronous speed.
In practice, this is enough to achieve almost all the advantages of variable-speed operation. This is the
most commonly used approach. A disadvantage is the maintenance requirements of the slip rings.
The wind power system design must optimize the annual energy capture at a given site. The only
operating mode for extracting the maximum energy is to vary the turbine speed with varying wind
speed such that at all times the TSR is continuously equal to that required for the maximum power
coefficient CP. The theory and field experience indicates that the variable speed operation yields 20 to
30 percent more energy than with the fixed speed operation. Nevertheless, cost of the variable-speed

| Power Control Pavesi - 174 -


control is added.
In the system design, this trade-off between the energy increase and the cost increase has to be
optimized. In the past, the added costs of designing the variable-pitch rotor, or the speed control with
power electronics, outweighed the benefit of the increased energy capture. However, falling prices of
the power electronics for speed control, and availability of high-strength fibre composites for
constructing high-speed rotors, have made it economical to capture more energy when the speed is
high. In the fixed-speed operation, on the other hand, the rotor is shut off during high-wind speeds,
losing significant energy.
The pros and cons of the variable and the fixed-speed operations are listed in Error! Reference
source not found..
We shall proceed by examining the variation of Pm as a function of ωm, with the wind speed U as
a parameter. The shaft power Pm for this turbine as a function of shaft rotational speed n is shown in
Figure 9.7.2. Pm is seen to rise to a maximum for each wind speed for a particular value of rotational
speed. Higher wind speeds have more power in the wind, and the change in tip speed ratio with
increasing wind speed causes the maximum to shift to a higher rotational speed. Maximum power is
reached at 38 r/min in a 6 m/s wind and
at 76 r/min in a 12 m/s wind. The
maximum possible shaft power in a 12
m/s wind is eight times that in a 6 m/s
wind, as would be expected from the
cubic variation of power with wind
speed.
It would appear that if fixed speed
operation is required, that 52.5 r/min is
a good choice for sites with a high
percentage of winds between 6 and 10
m/s since the turbine can deliver nearly
the maximum possible shaft power over
this wind speed range. A higher fixed

Table 9. 1 Advantages of Fixed and Variable Speed Systems


Fixed-Speed System Variable-Speed System
Higher rotor efficiency, hence, higher energy
Simple and inexpensive electrical system
capture per year
Fewer parts, hence higher reliability Low transient torque
Lower probability of excitation of mechanical
Fewer gear steps, hence inexpensive gear box
resonance of the structure
Mechanical damping system not needed, the
No frequency conversion, hence, no current
electrical system could provide damping if
harmonics present in the electrical system
required
Lower capital cost No synchronization problems
Stiff electrical controls can reduce system voltage
Figure 9.7.2 Shaft power output of Sandia 17-m Darrieus
sags in
variable-speed operation.

| Power Control Pavesi - 175 -


shaft speed would be justified only if there
were a significant fraction of wind speeds
above 10 m/s. Increasing the shaft speed
above 52.5 r/min would decrease the power
output for wind speeds below 8 m/s, and this
would be a significant penalty at many sites.
Suppose now that the load can accept
shaft power according to the curve marked
Load in Figure 9.7.2 and that the turbine is
free to operate at any speed. The turbine will
then operate at maximum power for any
wind speed, so the energy output will be
maximized. If this maximum energy output
can be obtained without increasing losses or
costs, then we have developed a system
which extracts more energy from the wind at
a lower cost per unit of energy.
Variable speed operation requires a
load which has a suitable curve of power
input versus rotational speed. The optimum
load will have a cubic variation of input
Figure 9.7.3 Shaft torque output of Sandia 17-m power versus rotational speed. In examining
Darrieus in variable-speed operation. various types of loads, we notice that the
input power to pumps and fans often has a
cubic variation with rotational speed. The power input to an electrical generator connected to a fixed
resistance will vary as the square of the rotational speed. The power input to a generator used to charge
batteries will vary even more rapidly than cubic. Load curves like the one shown in Error! Reference
source not found. are therefore quite possible, so variable speed operation is possible with several
applications.
In addition to the variation of shaft power with rotational speed, we must also examine the
variation of torque with rotational speed. If we take the shaft power curves in Figure 9.7.2 and divide
by the angular velocity at each point, we obtain the torque curves shown in Figure 9.7.3. One important
feature of these curves is that the torque is zero for n = 0, which means that the Darrieus turbine cannot
be expected to start without help. The friction in the transmission and generator produces an opposing
torque, and the torque Tm must exceed this opposing or tare torque before the turbine can be
accelerated by the power in the wind. The turbine must be accelerated to some minimum rotational
speed by an external power source so that Tm will reach a sufficient value to accelerate the turbine up to
operating speed. This particular machine would need to be accelerated up to perhaps 10 r/min before
the wind would be able to accelerate it to higher speeds.
This external power source is required for reliable starting of any Darrieus turbine. However,
there have been cases where a combination of very low tare torques and strong, gusty winds allowed a
Darrieus turbine to start on its own, sometimes with disastrous consequences.
A Darrieus turbine should have a brake engaged when repairs are being made or at other times
when rotation is not desirable.
The torque rises to a maximum at a particular rotational speed for each wind speed, in the same

| Power Control Pavesi - 176 -


manner as the shaft power. The peak torque is reached at a lower rotational speed than the peak shaft
power, as can be seen by a careful comparison of Figure 9.7.2 and Figure 9.7.3. The fact that maximum
shaft power varies as the cube of the rotational speed, we can argue that the maximum shaft torque
varies as the square of the rotational speed. In Figure 9.7.3, for example, the peak torque in a 12 m/s
wind is 10,600 Nm/rad at 60 r/min. The peak torque in a 6 m/s wind is 2650 Nm/rad at 30 r/min. The
peak torque has changed by a factor of four while the rotational speed has changed by a factor of two.
The turbine torque Tm must be opposed by an equal and opposite load torque TL for the turbine to
operate at a steady rotational speed. If Tm is greater than TL, the turbine will accelerate, while if Tm is
less than TL the turbine will decelerate. The mathematical relationship describing this is
d
Tm  TL  I
d
where I is the moment of inertia of the
turbine, transmission, and generator, all
referred to the turbine shaft.
The relationship between shaft torque
and an optimum load torque for the Sandia
17-m Darrieus turbine is illustrated in Figure
9.7.4. We have assumed a load torque with
the optimum variation
TL  Kn 2
The constant K is selected so the load
torque curve passes through the peaks of the
curves for turbine torque at each wind speed.
In order for the turbine to operate at
steady state or at a constant speed, the dωm/dt
term of Error! Reference source not found.
must be zero, and the load torque must be
equal to the shaft torque. Suppose that we
have a steady wind of 6 m/s and that the shaft
torque and load torque have reached
Figure 9.7.4 Load with square-law torque variation equilibrium at point a in Error! Reference
connected to Sandia 17-m turbine. source not found.. Now suppose that the
wind speed suddenly increases to 8 m/s and remains constant at that speed. The shaft torque Tm
increases to the value at a’ before the rotational speed has time to change. The load is still requiring the
torque at point a. Since the shaft torque is larger than the load torque, the turbine rotational speed will
increase until point b is reached, at which time the two torques are equal and steady state has again
been reached.
If the turbine is operating at point b and the wind speed suddenly decreases to 6 m/s, the turbine
torque will drop to the value at b’. Since the turbine torque is less than the load torque at this point, the
turbine will slow down until the turbine and load torques are again equal at point a.
The constant K of Error! Reference source not found. must be carefully selected for the
variable speed system to operate properly. Figure 9.7.5 shows the optimum load torque curve for the
Sandia 17-m turbine and also shows torque curves for two other loads where the load torque at a given

| Power Control Pavesi - 177 -


speed has been either doubled or halved. A
problem is immediately evident for the load
curve described by 2Kn2 in that the load torque
exceeds the turbine torque above a rotational
speed of about 16 r/min. The turbine will not
accelerate past this point and the system will be
characterized by a rather steady but very low
output. A load, such as a pump, with this torque
curve would have to be replaced with a load with
a smaller rating. Load torque specifications, as
well as turbine specifications, are often not
precisely known and would be expected to vary
slightly between two apparently identical
systems, so an adequate margin of safety needs
to be applied when selecting load sizes. It is
better to slightly undersize a load than to get it
so large that the turbine cannot accelerate to
rated speed.
The load specified as Kn2/2 requires
torques that are just slightly down the turbine
torque curve from the maximum torques. The
total energy production may be only 10 or 15
percent less for this load than that for the
optimum load. This reduction from the optimum
output may be acceptable in many applications,
Figure 9.7.5 Three loads attached to Sandia 17-m except for one major difficulty. The turbine
turbine rotational speed in a 12 m/s wind is 80 r/min for
the smaller load and 60 r/min for the optimum
load.
This greater speed may exceed the turbine speed rating and cause some speed limiting system to
be activated. We therefore need to have a load with a torque curve that falls between rather narrow
limits, perhaps within ten or twenty percent of the optimum torque curve. We shall examine torque
curves of specific loads in greater detail later in the text.
Turbines that are self-starting must have greater wind produced torques at low rotational speeds
than the Darrieus turbine. The torque at low rotational speeds will be proportional to the solidity of the
turbine, which is basically the fraction of the projected swept area that is actually covered by the rotor
blades. The Savonius turbine would have a solidity of 1.0, for example, with the American Multiblade
turbine approaching that value. Two-bladed horizontal axis turbines may have solidities closer to 0.1 at
the other extreme. The variation of torque with rotational speed for a low solidity, 10 m diameter, and
two bladed propeller is given in Figure 9.7.6. It may be noticed that the torque is greater than zero at
zero rotational speed, so the machine will start by itself. Once the machine is started, the torque curve
looks much like the torque curve for the Darrieus in that the torque increases with increasing rotational
speed until a peak torque is reached. The torque decreases rapidly past the peak until it reaches zero at
the run away or freewheeling rotational speed.
A high solidity machine may have its maximum torque at zero speed. This is the case for the

| Power Control Pavesi - 178 -


torque of the Savonius turbine shown in Figure
9.7.7. The high torque at low speeds may be
essential for some applications: for example,
starting a larger Darrieus or operating a positive
displacement pump.
We have shown that variable speed turbine
operation is technically possible. There is a
possibility of simplified construction and lower
costs and also a possibility of greater energy
production from a given rotor. Fixed speed
systems will probably dominate where induction
or synchronous generators are used to supply
power to a grid, but variable speed systems may
find a role in standalone situations as well as those
situations where mechanical power is needed
rather than electrical power.

9.7.3 Variable-slip operation


Variable slip represents a compromise
between fixed- and variable-speed operation
(Bossanyi and Gamble 1991, Pedersen 1995). The Figure 9.7.6 Torque versus rotational speed for
variable-slip generator is essentially an induction low-solidity 10-m-diameter two-bladed propeller.
generator with a variable resistor in series with the
rotor circuit, controlled by a high-frequency
semiconductor switch. Below rated, this acts just
like a conventional fixed-speed induction
generator. Above rated, however, control of the
resistance effectively allows the air-gap torque to
be controlled and the slip speed to vary, so the
behaviour is then similar to a variable-speed
system. A speed range of about 10 percent is
typical.
This is cheaper than a variable-speed system,
and gives some of the advantages, in particular the
control of torque in the drive train and the
smoothing of aerodynamic torque variations above
rated. It does not offer increased aerodynamic
efficiency below rated (although it does not suffer
from frequency converter losses), and it does not
allow any control of the power factor. Electrical
flicker will, however, be reduced above rated.
Slip rings can be avoided by mounting the Figure 9.7.7 Torque versus rotational speed for
variable resistors and control circuitry on the Savonius turbine: radius 0.88 m; area 22 m2.
generator rotor. An advantage of mounting these
externally via slip rings is that it is then easier to dissipate the extra heat which is generated above

| Power Control Pavesi - 179 -


rated, and which may otherwise be a limiting factor at large sizes.

9.7.4 Other approaches to variable-speed operation


There are other possible approaches to variable-speed operation, although none of these has found
significant commercial application. They include:
• use of a differential gearbox, with the third shaft controlled by a variable speed electric
motor/generator (Law, Doubt and Cooper, 1984, Burton, Mill and Simpson, 1990) or by a
hydraulic pump/motor (Henderson et al., 1990);
• mechanical continuously-variable transmission systems such as have been developed for
automotive applications.
Figure 9.7.8 depicts the distribution of the control methods used in small wind turbine designs.
Large machines generally use the power electronic speed control.

Figure 9.7.8 Speed control methods used in small to medium size


turbines.

| Power Control Pavesi - 180 -


9.8 References
American Gear Manufacturers Association/American Wind Energy Association, (1996).
‘Recommended practices for design and specification of gearboxes for wind turbine generator
systems’. AGMA/AWEA 921-A97.
American National Standards Institute/American Gear Manufacturers Association, (1995).
‘Fundamental rating factors and calculation methods for involute spur and helical gear teeth’.
ANSI/AGMA 2001-C95.
Anderson, C. G. et al., (1993). ‘Yaw system loads of HAWTS’. ETSU Report No W/42/00195/REP.
Energy Technology Support Unit, Harwell, UK.
Anderson, C. G., Heerkes, H. and Yemm, R., (1998). ‘Prevention of edgewise vibration on large stall-
regulated blades’. Proceedings of the BWEA Conference, pp 95–102.
Barbero, E. J., (1998). Introduction to composite materials design. Taylor and Francis, Philadelphia,
USA.
Bond, I. P. and Ansell, M. P., (1998). ‘Fatigue properties of jointed wood composites. Part I: Statistical
analysis, fatigue master curves and constant life diagrams’. J. Mat. Sci., 33,2751–2762, and ‘Part
II: Life prediction: analysis for variable amplitude loading’. J. Mat. Sci., 33, 4121–4129.
Bonfield, P. W. and Ansell, M. P., (1991). ‘Fatigue properties of wood in tension, compression and
shear’. J. Mat. Sci., 26, 4765–4773.
Bonfield, P. W. et al., (1992). ‘Fatigue testing of wood composites for aerogenerator blades. Part VII:
Alternative wood species and joints’. Proceedings of the BWEA Conference, pp 243–249.
Brinch Hansen, J., (1961). ‘The ultimate resistance of rigid piles against transverse forces’. Danish
Geotechnical Institute Report No. 12.
Brinch Hansen, J., (1970). ‘A revised and extended formula for bearing capacity’. Danish
Geotechnical Institute Bulletin No. 28.
British Standards Institution, (1986). ‘BS436: Spur and helical gears – Part 3 Method for calculation of
contact and root bending stress limitations for metallic involute gears’.
Corbet, D. C., (1991). ‘Investigation of materials and manufacturing methods for wind turbine blades’.
ETSU Report No. W/44/00261. Energy Technology Support Unit, Harwell, UK.
Corbet, D. C., Brown, C. and Jamieson, P., (1993). ‘The selection and cost of brakes for horizontal axis
stall-regulated wind turbines’. ETSU Report No. WN 6065. Energy Technology Support Unit,
Harwell, UK.
Echtermeyer, A. T., Hayman, E. and Ronold, K. O., (1996). ‘Comparison of fatigue curves for glass
composite laminates.’ Design of Composite Structures against Fatigue, (Ed. R. M. Mayer).
Mechanical Engineering Publications, Bury St Edmunds, UK.
European Convention for Constructional Steelwork, (1988). Recommendations on buckling of shells.
Freris, L. L. (ed) (1990). Wind energy conversion systems, Prentice Hall.
Fuglsang, P. L. and Madsen, H. A., (1995). ‘A design study of a 1 MW stall-regulated rotor’. Riso
National Laboratory Report No. R-799. Riso National Laboratory, Roskilde, Denmark.
Hancock, M. and Bond, I. P., (1995). ‘The new generation of wood composite wind turbine rotor
blades – design and verification’. Proceedings of the BWEA Conference, pp 47–52.
Hancock, M. Sonderby, O. and Schubert, M. A., (1997). ‘Design, development and testing of a 31 m
wood composite stall regulated blade for serial production’. Proceedings of the EWE Conference,
pp 206–212.
Heier, S., (1998). Grid integration of wind energy conversion systems. John Wiley & Sons, Chichester,

| Power Control Pavesi - 181 -


UK.
Hindmarsh, J., (1984). Electrical machines and their application. Pergamon Press, Oxford, UK.
Hu¨ ck, M., (1983). ‘Calculation of S/N curves for Steel, Cast Steel and Cast Iron – Synthetic S/N
curves’. Verein Deutsher Eisenhu¨ ttenleute Report No. ABF 11. Verlag Stahleisen, Du¨ sseldorf,
Germany.
Jamieson, P. and Brown, C. J., (1992). ‘The optimization of stall-regulated rotor design’. Proceedings
of the BWEA Conference, pp 79–84.
Jones, R. and Smith, G. A., (1993). ‘High quality mains power from variable-speed wind turbines’. IEE
Conference, Renewable Energy – Clean Power 2001.
Joose, P. A. and van Delft, D. R. V., (1996). ‘Has fatigue become a wearisome subject? –Overview of
12 years’ of materials research in the Netherlands’. Proceedings of the EUWEC Conference, pp
902–906.
Krause, P. C., (1986). Analysis of electric machinery. McGraw-Hill, New York, USA.
Mayer, R. M., (1996). Design of composite structures against fatigue. Mechanical Engineering
Publications, Bury St Edmunds, UK.
McPherson, G., (1990). An introduction to electrical machines and transformers. John Wiley & Sons,
New York, USA.
Mohan, N., Undeland, T. M. and Williams, W. P., (1995). Power electronics, converters application
and design, Second Edition. John Wiley & Sons, New York, USA.
Petersen, J. T. et al., (1998). ‘Prediction of dynamic loads and induced vibrations in stall’. Riso
National Laboratory Report No. R-1045. Riso National Laboratory, Roskilde, Denmark.
Pena, R., Clare, J. C. and Asher, G. M., (1996). ‘Doubly fed induction generator using back-to-back
PWM converters and its application to variable speed wind-energy generators’. IEE Proceedings
Electric Power Applications, 143, pp 231–241.
Schmidt, H. and Neuper, M., (1997). ‘Zum elastostatischen Tragverhalten exzentrisch gezogener L-
Sto¨ße mit vorgespannten Schrauben’ (‘On the elastostatic behaviour of an eccentrically
tensioned L-joint with prestressed bolts’), Stahlbau, 66, 163–168.
Schmidt, H., Winterstetter, T. A. and Kramer, M., (1999). ‘Non-linear elastic behaviour of imperfect,
eccentrically tensioned L-flange ring joints with prestressed bolts as basis for fatigue design’.
Proceedings of the European Conference On Computational Mechanics.
Thomsen, K., (1998). ‘The statistical variation of wind turbine fatigue loads.’ Riso National Laboratory
Report No. R-1063. Riso National Laboratory, Roskilde, Denmark.
Timoshenko, S. P. and Gere, J. M., (1961). Theory of Elastic Stability, Second Edition. McGraw-Hill,
New York, USA.
Van Delft, D. R. V., de Winkel, G. D. and Joose, P. A., (1996). ‘Fatigue behaviour of fibreglass wind
turbine blade material under variable amplitude loading’. Proceedings of the EUWEC
Conference, pp 914–918.
Verein Deutscher Ingenieure, (1986/1988). ‘VDI 2230 Part 1: Systematic calculation of high duty
bolted joints – Joints with one cylindrical bolt’.
Warneke, O., (1984). ‘Use of a double-fed induction machine in the Growian large wind energy
converter’. Siemens Power Engineering, VI, 1, pp 56–59.
Wilson, R. A., (1990). ‘Implementation and optimization of mechanical brakes and safety

| Power Control Pavesi - 182 -


10 Performance of wind energy conversion systems
For the efficient planning and successful implementation of any wind power project, an
understanding on the performance of the Wind Energy Conversion System (WECS), at the proposed
site, is essential. The major factors affecting the power produced by a WECS are (a) the strength of the
wind spectra prevailing at the site and its availability to the turbine (b) the aerodynamic efficiency of
the rotor in converting the power available in the wind to mechanical shaft power and (c) the
efficiencies in manipulating, transmitting and transforming this power into the desired form. Hence,
assessment of the performance of a WECS is rather a complex process.
Wind is a stochastic phenomenon. Velocity and direction of wind at a location considerably vary
from season to season and even time to time, as we have discussed in chapter 4. Hence, apart from the
strength of the wind spectra, its distribution also has significant influence on the performance of the
WECS. Further, the characteristic operational speeds of a wind machine should match well with the
prevailing wind spectra to ensure the maximum exploitation of the available energy. In this chapter, we
will define the performance of different WECS by integrating these site parameters with the machine
characteristics.
Several types of wind machines with the same rated capacity but different operating
characteristics may be available in the market. The designer of a wind energy project should
judiciously choose the turbine optimally matching with his site to maximise the energy production.
Hence, the criteria for selecting the ‘right machine for the right site’ will also have to be included in our
discussions.
As in Chapter 4, most of the expressions presented here cannot be solved through straightforward
analytical methods.

10.1 Power curve of the wind turbine


One of the major factors affecting the performance of a WECS is its power response to different
wind velocities. This is usually given by the power curve of the turbine. The power curve of the
machine reflects the aerodynamic, transmission and generation efficiencies of the system in an
integrated form.
Figure 10.1.1 shows the typical power curve of a pitch controlled wind turbine. The rated power
of the turbine is 1 MW. The given curve is a theoretical one and in practice we may observe the
velocity power variation in a rather scattered pattern.
We can see that the important characteristic speeds of the turbine are its cut-in velocity (VI), rated
velocity (VR) and the cut-out velocity (VO). The cut-in velocity of a turbine is the minimum wind
velocity at which the system begins to produce power. It should not be confused with the start-up speed
at which the rotor starts its rotation. The cut-in velocity varies from turbine to turbine, depending on its
design features. However, in general, most of the commercial wind turbines cut-in at velocities
between 3 to 5 m/s.
Due to technical and economical reasons, the wind turbine is designed to produce constant power
- termed as the rated power (PR) - beyond its rated velocity .
Thus, the rated velocity of a turbine is the lowest wind velocity corresponding to its rated power.
Usually the system efficiency is maximum at VR. From VI, to VR, the power generated by the turbine
increases with the wind velocity. Between VR and VO, the turbine is restricted to produce constant
power PR corresponding to VR, irrespective of the changes in velocity. This power regulation is for
better system control and safety. Hence PR is the theoretical maximum power expected from the

| Performance of wind energy conversion systems Pavesi - 183 -


Figure 10.1.1 Ideal power curve of a pitch controlled wind turbine
turbine. At wind velocities higher than VO, the machine is completely shut down to protect the rotor
and drive trains from damage due to excessive loading. Sometimes VO is also termed as the furling
velocity, as in the earlier sail type machines the canvas was rolled up to protect the mill from strong
winds.

Table 11. 1 Performance regions of a wind turbine


Velocity range Power
0 to VI No power as the system is idle
VI to VR Power increases with V
VR to VO Constant power PR
Greater than VO No power as the system is shut down

Hence, the turbine has four distinct performance regions as indicated in Table 11. 1. The power
produced by the system is effectively derived from performance regions corresponding to V I to VR and
VR to VO. Let us name these as region 1 and 2. The velocity-power relationship in the region 1 can be
expressed in the general form
P  aV n  b (10.3.4)
where a and b are constants and n is the velocity-power proportionality. Now consider the performance
of the system at VI and VR. At VI, the power developed by the turbine is zero. Thus
aVIn  b  0 (10.3.5)
At VR power generated is PR. That is:

| Performance of wind energy conversion systems Pavesi - 184 -


PR  aVRn  b (10.3.6)
Solving (10.3.4) and (10.3.6) for a and b and substituting in (10.3.6) yields
 Vn  VIn 
PV  PR  n n 
(10.3.7)
 VR  VI 
Equation (10.3.7) gives us the power
response of the turbine at wind velocities
falling under the region 1. Obviously, the
power corresponding to the region 2 is PR.
In practice, the power response of
the turbine does not follow a perfect curve
as shown above. Field observations are
rather scattered (Figure 10.1.2). This
means that, for the same wind velocity V,
the turbine appears to develop different
powers as shown in the figure. This
happens because we cannot precisely
measure the strength of the wind as it
passes through the turbine. The
measurements are made from
meteorological towers mounted slightly
Figure 10.1.2 Field performance of a 250 kW wind turbine away from the turbine. Due to momentary
fluctuations in the wind, velocity of the
measured wind and the wind column
actually interacting with the turbine may
be different. Further, due to the variations
in the air density, there may be differences
in the power characteristics recorded
during different days. Observations from
wind farms indicate that even the same
type of turbines installed at the same
location may show different power
characteristics [Li S et al. (2001)]. It
should also be noted that the power curve
tells us the power expected from the
turbine at an instantaneous wind velocity
V and not for the velocity averaged over a
period.
The velocity power proportionality
has a profound effect on the power
generated by the wind turbine. Ideally,
variations in power with the velocity
Figure 10.1.3 Correlation between measured and generated should be cubic in nature.
performances However in practice, this can take

| Performance of wind energy conversion systems Pavesi - 185 -


any form such as linear, quadratic, cubic or even higher powers and its combinations [Lysen, E.
H.,(1983), Mathew S (2002)]. For example, let us take the performance of the turbine considered in
Figure 10.1.2.
Velocity-power characteristics of this turbine for different n are simulated. The correlation
coefficients between the observed and generated performances are plotted in Figure 10.1.3. The
correlation is highest for n = 2. Thus, this turbine shows a quadratic velocity-power relationship in its
1st performance region. Value of n varies from turbine to turbine. For Darrieus turbine, the velocity
power relationship is generally linear [Johnson GL (2001)], where as some two bladed horizontal axis
designs show cubic velocity-power response.

10.2 Energy generated by the wind turbine


The factors influencing the energy produced by a WECS at a given location over a period are: (1)
the power response of the turbine to different wind velocities (2) the strength of the prevailing wind
regime and (3) the distribution of wind velocity within the regime. Energy corresponding to a certain
wind velocity V is the product of the power developed by the turbine at V and the time for which this
velocity V prevails at the site. The total energy generated by the turbine over a period can be computed
by adding up the energy corresponding to all possible wind speeds in the regime, at which the system is
operational. Hence, along with the power characteristics of the turbine, the probability density
corresponding to different wind speeds also comes into our energy calculations.
This is expressed graphically in Figure 10.2.1. In the figure, curve (a) represents the power curve
of a commercial 1 MW turbine with 3.5 m/s cut-in, 15 m/s rated and 25 m/s cut-out velocities. Curve
(b) represents the probability density function at a site with 8 m/s mean wind speed, assuming the
Rayleigh distribution. The energy corresponding to different wind velocities are plotted in curve (c).
To find out the energy contributed by a wind speed of 10 m/s, the power and the probability
density corresponding to 10 m/s are located from the curves (a) and (b). The energy corresponding to
this velocity is indicated in the curve (c), which is basically the product of the power and the
probability density for 10 m/s. Total energy available from this turbine at this site, over a period, can be
estimated by integrating curve (c) within the limits of the cut-in and cut-out velocities and multiplying
it with the time factor. Let us express these in mathematical terms in the following sections, adopting
both the Weibull as well as Rayleigh distributions.

10.2.1 Weibull based approach


As we have seen in Chapter 4.6, if the wind velocity in a regime follows the Weibull distribution,
its probability density function f(V) and cumulative distribution function F(V) are given by
 c
k

FV  1 e
 V
(10.4.1)

and
k 1
dF k  V   c
k

f V 
 V
   e (10.4.2)
dV c  c 
Here k is the Weibull shape factor and c is the scale factor. Methods to determine k and c are
described in Chapter 4.6.2.1.

| Performance of wind energy conversion systems Pavesi - 186 -


Figure 10.2.1 Graphical representation of energy estimation
Referring to Figure 10.2.1, let EIR and ERO be the energy developed by the system at its
performance regions 1 and 2 respectively. Then EIR can be computed by adding the energy
corresponding to all possible wind velocities between VI and VR. Thus,
VR

E IR  T  PV f  V  dV (10.4.3)
VI

Similarly, ERO can be expressed as


VO

E RO  TPR  f  V  dV
VR
(10.4.4)

Substituting for P(V) and f(V) from Eqs. (10.3.7) and (10.4.2) in Eq. (10.4.3), we get

| Performance of wind energy conversion systems Pavesi - 187 -


k 1
 V n  VIn kV  
VR k
 V
E IR  PR T   n    e c dV
V  VIn
VI  R c c 
k 1
(10.4.5)
 PT  R  
V
kV
  n R n    V n  VIn   
k
 V
e c dV
 VR  VI  VI c c 

Now, let us introduce the variable X such that


k
V
X  (10.4.6)
c
Then,
k 1
kV
dX   
c c 
and (10.4.7)
1
V  cX k

With Eq. (10.4.6), we have


k
V 
XI   I 
 c 
k
V 
XR   R 
 c  (10.4.8)

k
V 
XO   O 
 c 
Thus, after simplification, EIR can be expressed as
X
PR Tcn PR TVIn
 
R


n
EIR  X k X
e dX  e  XI  e  XR (10.4.9)
 VR  VI  XI
n n
 VR  VI 
n n

Now consider the second performance region. From Eq. (10.4.4), ERO may be represented as
k 1
 c  dV
VO
kV  V
k

E RO  PR T    e (10.4.10)
VR
c c 
As

 f  V  dV  F  V  (10.4.11)

from Eqs. (10.3.7) and (10.4.1), Eq. (10.4.10) can be simplified as


E RO  Pr T e Xr  e Xo  (10.4.12)

| Performance of wind energy conversion systems Pavesi - 188 -


The total energy generated by the wind turbine, over a given period of time, is the sum of energy
derived from the performance regions 1 and 2. Thus,
E T  E IR  E RO (10.4.13)

10.2.2 Rayleigh based approach


Rayleigh distribution is a simplified case of the Weibull distribution which is derived by
assuming the shape factor as 2. Owing to its simplicity, this distribution is widely used for wind energy
modeling. Under the Rayleigh based approach, the cumulative distribution and probability density
functions of wind velocity are given by
 2
  V 

FV  1 e 4  Vm 
(10.4.14)
and
 2
 V  4  V Vm 
f V  e (10.4.15)
2 Vm2
As in the previous case, let us consider the performance regions 1 and 2 of the power curve
(Figure 10.2.1). Referring to Eq. ((10.4.3)), energy developed by the turbine (EIR) , while VI < V < VR,
may be represented as
 2
 V  4  V Vm 
VR
PT
E IR  n R n
VR  VI   Vn
 V  I
n

2 Vm2
e dV (10.4.16)
VI

Substituting
2
 V 
X  
4  Vm 
 V
dX  dV
2 Vm2 (10.4.17)
and
1
 X 2
V  2Vm  

Eq. (10.4.17) implies that

| Performance of wind energy conversion systems Pavesi - 189 -


2
 V 
XI   I 
4  Vm 
2
 V 
XR   R 
4  Vm  (10.4.18)
and
2
 V 
XO   O 
4  Vm 
Thus we can express EIR as
XR X
2n Vmn PR T PR TVIn R

 
n
2 X
E IR  n X e dX  e X dX (10.4.19)
  VR  VI  XI
2 N N
 VR  VI  XI
N N

This is further simplified as


XR
2n Vmn PR T

PR TVIn
 
n
E IR  X 2 e X dX  e  XR  e  XI (10.4.20)

n
2
 VRN  VIN  XI  VR  VI 
N N

Now considering the performance of the system while VR< V< VO, we get
2
 V  4 V Vm 
VO

E RO  TPR 
VR
2 Vm2
e dV (10.4.21)

With  f  V  dV  F  V  the above expression can further be simplified as


XO

e
X
E RO  T PR dX
XR (10.4.22)

 T PR e XR  e XO 
Once we have expressions for EIR and ERO, the total energy developed by the system ET is
obtained by adding EIR and ERO.

10.3 Capacity factor


Capacity factor is one of the important indices for assessing the field performance of a wind
turbine. The capacity factor (CF) of a WECS at a given site is defined as the ratio of the energy actually
produced by the system to the energy that could have been produced by it, if the machine would have
operated at its rated power throughout the time period. Thus
ET
CF  (10.5.1)
T PR
The capacity factor reflects how effectively the turbine could harness the energy available in the

| Performance of wind energy conversion systems Pavesi - 190 -


wind spectra. Hence, CF is a function of the turbine as well as the wind regime characteristics. Usually
the capacity factor is expressed on an annual basis. Capacity factor for a reasonably efficient turbine at
a potential site may range from 0.25 to 0.4. A capacity factor of 0.4 or higher indicates that the system
is interacting with the regime very efficiently.
Information on the capacity factor of the turbine at a given site may not readily available during
the initial phases of project identification. Under such situations, it is advisable to calculate the rough
capacity factor (RCF). This is basically deduced from the power curve of the machine, based on the
average wind velocity at the site. From the power curve, we can locate the power corresponding to the
average wind velocity. Dividing this power (PVm) by the rated power of the turbine, the rough capacity
factor can be calculated. Thus,
PVm
R CF  (10.5.2)
PR
This is a very crude method. Wind energy potential of the site should be analysed in detail,
following the methods discussed in the previous sections, before we proceed further with the project. It
should be noted that, a system offering the highest capacity factor may not be always the most
economic choice for a location.
For example, consider a site having very strong wind spectra. Two wind turbines of the same
rotor area, but different in generator size is being installed at this site. The turbine with smaller
generator may show higher capacity factor at the site. However, the second system with the bigger
generator may produce more energy at the site, in spite of its lower capacity factor. Thus the second
turbine may prove to be more economical.

10.4 Matching the turbine with wind regime


By now, it is evident that performance of a WECS at a site depends heavily on the efficiency with
which the turbine interacts with the wind regime. Hence, it is essential that the characteristics of the
turbine and the wind regime at which it works should be properly matched. The capacity factor of the
system can be a useful indication for the effective matching of wind turbine and regime. For turbines
with the same rotor size, rated power and conversion efficiency, the capacity factor is influenced by the
availability of the turbine to the prevailing wind.
In other words, the functional velocities of the turbine (VI, VR and VO) should be chosen in such a
way that, the energy available with the wind regime is exploited to its maximum level. This in turn
would require that the turbines should be individually designed for each site so that these functional
parameters can be defined according to the site characteristics. This is not practical.
Several wind turbines of different ratings and functional velocities are available in the market. A
wind energy project planner can choose a system, best suited for his site, from these available options.
Hence, it is important for him to identify the effect of these functional velocities- that is VI, VR and VO
- on the turbine performance at the given location and ensure that the turbine and the wind regime are
working in harmony. The performance estimation methods discussed in the above sections can be used
for such an analysis.
VO is kept at a reasonably higher value for this analysis. The capacity factors thus obtained are
plotted against the respective cut-in speeds in Figure 10.4.1. Up to a velocity of 3.0 m/s, the cut-in
velocity does not have any significant influence on the capacity factor. However, for cut-in velocities
higher than 3.0 m/s, there is a noticeable decrease in the capacity factor.
Similar procedure may be followed for identifying optimum VO of the turbine. In this case, VI is

| Performance of wind energy conversion systems Pavesi - 191 -


fixed and VO is varied. Results
are shown in Figure 10.4.1.
Effect of VO on the system
performance is prominent up to
16 m/s. With further increase in
the cut-out velocity, the capacity
factor is not improved
considerably.
In the above procedure, we
have identified the effect of V I
and VO based on the site
characteristics. However, the
machine characteristics also have
to be considered while choosing
VI and VO for a system. The cut-
in speed should be strong enough
to overcome all the system
losses. Considering the
efficiencies of power
transmission and generation, a
higher cut-in speed, say 0.3 to
Figure 10.4.1 Effect of cut-out velocity on the system performance. 0.5 times the rated speed, may be
suggested in some situations
[Pallabazzer R (1995)]. Similarly, capability of the system in sustaining extreme aerodynamic loads
should be considered while fixing up VO. Owing to engineering and economic reasons, the cut-out
speed normally does not exceed
2VR in most of the commercial
designs.
This indicates that VI and VO
are influenced by VR. For a given
rotor area and efficiency, the rated
speed is directly correlated with the
system’s rated power.
Now let us examine the effect
of VR on the capacity factor taking
VI = 3 m/s VO = 16 m/s (Figure
10.4.3). The capacity factor
decreases with increase in VR.
The reason is evident from
Figure 10.4.3, in which the power
curves of two similar systems,
which differ only in the rated
velocity, are compared. With the
increase in VR, area under the
power curve reduces, which is
Figure 10.4.2 Effect of cut-in velocity on the system performance

| Performance of wind energy conversion systems Pavesi - 192 -


finally reflected as the reduction
in the capacity factor. However,
for the same rated power, lower
rated velocity will in turn demand
a bigger rotor for the turbine. As
the capital investment required for
the WECS is proportional to the
rotor area, this will increase the
unit cost of energy produced.
For a given rotor size,
increase in the rated velocity
means increase in the rated power
and thus the generator size
(assuming that the efficiency is
unchanged). If this higher rated
velocity is justified by the
strength and nature of the
prevailing wind regime, this
would in turn improve the energy
production and thus reduce the
Figure 10.4.4 Capacity factor as a function of rated wind velocity cost of unit energy generated. At
the same time, if the rated wind
speed is too high for the regime, the system will seldom function at its rated capacity.
The above discussions are
hypothetical and meant only to
demonstrate the effect of VI, VR and VO
on the turbine performance. Unless under
special situations, it is not practical to
design wind turbines for a specific site.
So let us look into this problem in a more
practical point of view.
Several wind machines of the same
power class but differing in performance
curves may be commercially available.
Designer of the wind energy project often
chooses a system from these available
options for his site. Selection of the ‘right
machine for the right site’ plays a major
role in the success of the project.
Depending on the Weibull scale and
shape factors, it is possible to identify VR
suitable for a particular wind regime
[Galanis N, Christophides C (1990),
Jangamshetti SH, Rau VG (2001),
Figure 10.4.3 Comparison of power curves for different Pallabazzer R (1995)]. It is suggested that
rated wind speeds.

| Performance of wind energy conversion systems Pavesi - 193 -


the rated wind speed may be about twice the average wind speed for regimes with k = 2. In trade winds
with higher k, VR may be 1.3 times Vm.

10.5 Performance of wind powered pumping systems


One of the potential applications of wind energy is for water pumping, especially in remote rural
areas, where grid connected power supply is not readily available.
Several thousands of such systems are still functional in different parts of the world. These pumps
can have either mechanical or electrical power transmission.
Performance of a wind pump in a given wind regime is governed by (1) the behaviour of its two
major components viz. the wind rotor and the pump. (2) Effective interaction of the rotor-pump
combination with the wind regime. Several methods are proposed to estimate the wind pump
performance under fluctuating conditions of wind regimes. In the most common approach, a constant
overall efficiency is assumed throughout the working range of the system. Here, the pump and wind
regime characteristics are not considered. Though this sounds simple and handy, this approach may
result in overestimating the capacity.
Some elaborate methods, taking the rotor and the pump characteristics into account, were also
suggested [Biswas S, et al (1994), Powel WR (1981)]. However, the interaction of the rotor-pump
combination with the wind regime is overlooked in these models. Hence, these models give only the
instantaneous performance of the system in terms of expected discharge at a given wind velocity.
We will follow an integrated approach incorporating the characteristics of the rotor, pump and the
wind regime, for defining the system performance. All the common types of systems viz. (1) wind
driven piston pumps with mechanical transmission, (2) wind driven roto-dynamic pumps with
mechanical transmission and (3) wind driven roto-dynamic pumps with electrical transmission, are
considered.
Most of the wind pumps are to be installed in remote rural areas. Wind data available for such
locations are in the form of seasonal averages (for example daily or even monthly mean wind velocity).
Hence Rayleigh distribution is used for our analysis in the following sections.

10.5.1 Wind driven piston pumps


A simple multi-bladed windmill mechanically coupled to a piston pump without any back gearing
and power regulation is considered here. Such pumps are popular in many parts of the world, especially
in developing countries.
These pumps are designed to work at low wind velocity regions. Some of these systems have V I
as low as 2.5 m/s and VO around 10 m/s. The output increases steadily from VI to VO, without any
region of constant power as shown in Figure 10.5.1.

10.5.2 Rotor performance


Usually, the rotors used for water pumping wind mills show quadratic velocity power
relationship. Hence, we can express the power PV at any wind velocity V as
PV  aV 2  b (10.7.1)
where a and b are constants. At the cut-in condition, the power is zero and thus
aVI2  b  0 (10.7.2)

| Performance of wind energy conversion systems Pavesi - 194 -


The system reaches at its maximum power at the cut-out point. Hence the power PO,
corresponding to VO can be represented as
PO  aVO2  b (10.7.3)
Solving Eqs. (10.7.2) and (10.7.3) for a and b and substituting in Eq. (10.7.1), we get

 V2  VI2 
PV  PO  2 2 
(10.7.4)
 VO  VI 
The above equation is similar to the expression that we have developed for the wind generators.
Test result of an eight bladed-5 m diameter experimental rotor developed for a water pumping wind
mill are shown in Figure 10.5.1. VI and VO of the rotor are 2.5 and 10 m/s respectively. Performance of
the rotor generated using Eq. (10.7.4) is also indicated in the figure as continuous line. Reasonable
agreement can be seen between the estimated and observed performances.
For estimating the energy produced by the rotor at a given site over a period, we have to introduce
the wind regime characteristics in our analysis. The energy developed (EI) over the period T is given by
VO

EI  T  PV f  V  dV (10.7.5)
VI

Assuming the Rayleigh distribution for wind velocity and substituting for f(V) and P(V), we get
 V2  VI2   V  4 V Vm 
VO 2

E I  T  PO  2 2 
e dV (10.7.6)
 VO  VI  2 Vm
2
VI

Substituting
2
 V 
X  
4  Vm 
2
 V 
XI   I 
4  Vm  (10.7.7)
and
2
 V 
XO   O 
4  Vm 
XO
TP 4Vm2
EI  2 O 2   X  X I  e X dX
VO  VI XI

(10.7.8)
T P 4Vm2
  X  X I  e X  e X
XO
 2 O 2
VO  VI  XI

Further simplification yields

| Performance of wind energy conversion systems Pavesi - 195 -


 
E I  T PO 
4Vm2
   VO  VI 
2 2
e XI
 e  XO

 e  XO


(10.7.9)

Figure 10.5.1 Typical power curve of a multi-bladed mechanical wind pump

Figure 10.5.2 Field performance of the experimental rotor

| Performance of wind energy conversion systems Pavesi - 196 -


10.5.3 Wind pump characteristics
The load imposed on the turbine by a positive displacement pump is typical in nature.
For an unbalanced single acting pump, the starting torque is π times higher than the average
running torque [Burton JD (1983)]. This means that, once the system stops due to the fluctuations in
the wind regime, it restarts only at a wind speed strong enough to produce this high torque. Hence,
while estimating the performance of such a wind pump, apart from the rotor and wind regime
characteristics, the pump behaviour should also be taken into account. The overall performance
coefficient of a wind rotor coupled to a reciprocating pump can be modelled as [Burton JD, Pinilla AE
(1985), Mathew S, Pandey KP (2000)]
 V  
2
V 
2

CP   4CPd  T,P  1  K O  I   K O  I  (10.7.10)


  V   V

where CP  is the overall (wind to water) efficiency of the system, CPd is the power coefficient of
the rotor at the design point, (T,P) is the combined transmission and pump efficiency and K0 is a
constant taking care of the starting behaviour of the rotor pump combination. For a single acting pump,
K0 varies from 0.2 to 0.25.
The power developed by the system in pumping water (PV) is given by
1
PV  CP  a AV3 (10.7.11)
2
From Eqs. (10.7.10) and (10.7.11), PV can be expressed as
  VI  
2
 VI 
2

PV  2CPd  T,P a AV 1  K O    K O  


3
(10.7.12)
  V   V

Figure 10.5.3 compares the field performances of a commercial wind pump, with the performance
computed using Eq. (10.7.12). Close agreement can be observed between the measured and calculated
performances.
The hydraulic power (PH), needed by the pump for delivering a discharge of QVP against a head h
is given by
PH  w gQVP h (10.7.13)
Equating the power delivered by the rotor and absorbed by the pump, and solving for Q VP, the
instantaneous discharge of the system at any velocity V can be derived.
Thus,

    AV3   
2 2
V   VI 
QVP  2CPd  T,P   a    1  K O  I   KO   (10.7.14)
 w   gh   V  V

10.5.4 Integrated system performance


In order to estimate the discharge of the wind pump over a period, let us integrate the rotor, pump
and wind regime characteristics into our model. In light of Eq. (10.7.12), the power PO corresponding
to the velocity VO can be rewritten as

| Performance of wind energy conversion systems Pavesi - 197 -


  VI  
2
 VI 
2

PO  2a AV CPd  T,P  1  K O    K O  


3
O (10.7.15)
  V   V
Similarly, the energy required to pump a total discharge of QIP over a period is given as
E H  w gQIP h (10.7.16)
Substituting for PO in the expression for EI in Eq. (10.7.9), equating it with hydraulic energy
demand EH , and solving for QIP, we get

 a   AV3    VI    VI   
2 2

QIP  2TCPd  T,P      1  K O   K O  


4Vm2
 V     VO2  VI2 
e 
 XI
 e  XO

 e  XO

 w   gh    V   
(10.7.17)
Thus Eq. (10.7.17) gives us the discharge expected from a wind driven piston pump, installed at a
given site, over a period T.

10.5.5 Wind driven roto-dynamic pumps


Wind turbines mechanically coupled with roto-dynamic pumps are increasingly getting attention
in the recent years. These pumps are suitable for pumping large quantities of water from shallow wells
and ponds. Let us consider the performance of such a system under fluctuating conditions of wind
regimes. Here we assume that the efficiency of the combined system peaks at the operating point where
the efficiency of the pump is also at its maximum.
The power (PTd) developed by a wind turbine at its design wind speed (Vd) is given by
1  D2 
PTd  CPda  T  Vd3 (10.7.18)
2  4 
Where CPd is the design power coefficient and DT is the rotor diameter. Considering the
characteristics of an ideal roto-dynamic pump, the power required (Ppd) for delivering its design
discharge (Qd) against the design head (hd) may be expressed as
w gh d Qd
PPd  (10.7.19)
Pd
where Pd is the design efficiency of the pump.
Assuming that both the rotor and the pump are matched at their peak efficiency points, we can
equate the power developed by the turbine with that required by the pump. Thus, discharge of the
combined system at its design point can be solved as
1   DT2  Vd3
Qd  CPd Pd T a   (10.7.20)
2 w  4  gh d
Here T is the efficiency of power transmission from the wind rotor to the pump.
As the wind pump is exposed to fluctuating conditions of wind regime, the off design
performance of the system is also important. For an ideal roto-dynamic pump, the discharge (Q), pump
speed (NP) and the pump diameter (Dp) can be correlated as

| Performance of wind energy conversion systems Pavesi - 198 -


Figure 10.5.3 Field performance of wind driven piston pumps
Q  N p D3p (10.7.21)
Thus, the discharge of the wind pump at any velocity V, may be expressed in terms of its design
discharge as
N 
QV  Qd  pv  (10.7.22)
 Npd 
 
where QV and Npv are the discharge and pump speed at velocity V and Npd is the design pump speed.
As the pump, at all its operating speeds, is matched with the best efficiency points of the turbine,
the rotational speed of the pump at any velocity can be represented by
 1 
Npv  Gd V   (10.7.23)
 DT 
where G is the gear ratio to be introduced between the wind rotor and pump and d is the design tip
speed ratio of the rotor. From Eqs. (10.7.20), (10.7.22) and (10.7.23), the pump discharge at any wind
velocity can be expressed as

1  a  Vd3   Gd 
QV  Cpd pd T VDT      (10.7.24)
8  w  gh d   Npd 

| Performance of wind energy conversion systems Pavesi - 199 -


Eq. (10.7.24) gives us the system discharge against the design head. However, due to changing
field conditions, the lift requirement also may vary. As for the same power discharge is inversely
proportional to head, we have
hd
Q V,h   QV (10.7.25)
h
where Q(V,h) is the discharge of the system at a velocity V against a head h. Combining Eqs. (10.7.24)
and (10.7.25), we can express Q(V,h) as

1    V3   G 
Q V,h   Cpd pd T VDT  a  d   d  (10.7.26)
 
8  w   gh   Npd 
If Npi be the rotational speed of the pump required to start pumping against a given head, then
N pi
NT  (10.7.27)
G
Hence, the cut-in velocity of the wind driven roto-dynamic pump at a certain hydraulic load can
be expressed by
DT N pi
VI  (10.7.28)
G d

10.5.6 Integration of wind regime characteristics


Let us now introduce the wind regime characteristics in the analysis to quantify the pump
discharge, at a given site, over a period. Here we assume that the system cuts in at VI and cuts out at VO
without any region of constant power. With this assumption, discharge of the wind pump over a period
(QIR) is given by [Mathew S, Pandey KP (2003)]
VO

QIR  T  Q V.h f  V  dV (10.7.29)


VI

Assuming the Rayleigh distribution for the wind velocity and with
2
 V 
X   (10.7.30)
4  Vm 
QIR can be represented as

  a  Vd3  1   Gd  O 2  X
V

QIR  TDT Cpd pd T V   2   


   V e dX (10.7.31)
16  w   Vm   gh   Npd  VI

10.5.7 Wind electric pumping systems


Wind pumps with electrical transmission are more efficient, reliable and flexible compared to the
mechanical systems. These pumps have the performance regions similar to that of the wind electric
generators described in section 10.2 Detailed performance models for wind electric pumps with

| Performance of wind energy conversion systems Pavesi - 200 -


variable speed permanent magnet generators and induction motor drives are discussed in [Mathew S, et
al. (2002)]. Here, we adopt a simpler method to approximate the performance of these pumps.
Referring to Error! Reference source not found., in the first performance region we have
w ghQ  Vn  VIn 
 PR  n n 
(10.7.32)
  VR  VI 
From the above expression, discharge of the pump (QV) at velocity V and head h can be solved as
  Vn  VIn 
QV  PR   (10.7.33)
w gh  VRn  VIn 
where  is the pump efficiency. For the second region of constant power,

QV  PR (10.7.34)
w gh
To estimate the integrated discharge from these pumps over a period, the energy developed by the
system may be computed as discussed in case of the wind generators and then equated with the
corresponding hydraulic energy demand. Thus

QIE  EI (10.7.35)
w gh

| Performance of wind energy conversion systems Pavesi - 201 -


10.6 References
Biswas S, Sreedhar BN, Singh YP (1994) A simplified statistical technique for wind energy output
estimation. Wind Eng. 15 (10) : 921-930
Burton JD (1983) Double acting pump with inertia flow improves the load matching of water pumping
windmills. In: Seventh Conference of Fluid Machinery, Budapest
Burton JD, Pinilla AE (1985) Water pump for wind mills – A comparison between two commercially
available systems from South America. Wind Eng. 9 (1) : 50-58
Galanis N, Christophides C (1990) Technical and economical considerations for the design of optimum
wind energy conversion systems. J Wind Eng. Ind. Aerodynamics 34 : 185-196
Jangamshetti SH, Rau VG (2001) Optimum siting of wind turbine generators. IEEE Transactions on
Energy Conservation 16 (1) : 8-13
Johnson GL (2001) Wind energy systems. http://www.rpc.com.au
Li S, Wunsch DC, O’Hair EA, Giesselmann MG (2001) Using neural networks to estimate wind
turbine power generation. IEEE Transactions on Energy Conversion 16 (3) : 276-282
Lysen, E. H.,1983. Introduction to Wind Energy. CWD, Amersfoort, The Netherlands
Mathew S, Pandey KP (2000) Modelling the integrated output of mechanical wind pumps. J Solar
Energy Engineering 122
Mathew S, Pandey KP (2003) Modelling the integrated output of winddriven roto-dynamic pumps.
Renewable Energy 28 : 1143-1155
Mathew S, Pandey KP, Burton JD (2002) The wind-driven regenerative water-pump. Wind engineering
26(5) : 301-313
Mathew S, Pandey KP, Kumar AV (2002) Analysis of wind regimes for energy estimation. Renewable
Energy 25: 381-399
Pallabazzer R (1995) Evaluation of wind – generator potentiality. Solar Energy 55 (1) : 49-59
Powel WR (1981) An analytical expression for the average output power of a wind machine. Sol
Energy 26 : 77-80

| Performance of wind energy conversion systems Pavesi - 202 -


11 Environmental Aspects and Impacts
This chapter will review the environmental aspects associated with the deployment of a single
wind turbine or a wind farm. Wind energy development has both positive and negative environmental
impacts. On the positive side, wind energy is generally regarded as environmentally friendly, especially
when the environmental effects of emissions from large-scale conventional electrical generation power
plants are considered. For example, estimates of the stack emissions (sulphur and nitrogen oxides,
particulates, and carbon dioxide) of conventional coal and gas-fired power plants as compared to those
of wind systems (zero in all cases) are shown in Table 13. 1.

Table 13. 1 Stack emissions of coal, gas, and wind power plants (kg/MWh)

Pollutant Conventional coal w/ controls Conventional gas w/controls Wind


Sulphur oxides 1.2 0.004 0
Nitrogen oxides 2.3 0.002 0
Particulates 0.8 0.0 0
Carbon dioxide 865 650 0

Table 13. 2 Annual reduction in atmospheric emissions due to a 5 MW


wind farm
Pollutant Reduction, kg
Carbon dioxide 8932178
Nitrous oxide 645.6996
Sulphur dioxide 6456.996
NOX 22599.49
Particulates 1076.166

The stack (or direct) emissions of wind systems are essentially zero, although there are indirect
emissions associated with the actual production of wind turbines and the erection and construction of
wind turbine systems and wind farms. Estimates of indirect emissions from wind systems (in Germany)
are presented in the review paper of Ackerman and Soder (2000), where the emissions values are
shown to be generally small (one or two orders of magnitude less than those of conventional power
plants).
The determination of the overall cost to society of the various emission pollutants is difficult to
measure and open to much debate. In general, the environmental benefits of wind power are calculated
by the avoided emissions from other sources. These benefits generally come from the displacement of
generation (MWh) as opposed to capacity (MW).
The displacement of capacity does have its benefits, however, especially if extension of the fuel
and water supply infrastructure can be avoided (Connors, 1996). As more wind turbines and wind farm
systems have been planned or installed in the United States and Europe, the importance of their
environmental effects has increased. This is well documented in many publications on the positive
environmental aspects of wind power. However, some negative environmental aspects have been
shown to be especially important in populated or scenic areas. Here, a number of potential wind energy

| Environmental Aspects and Impacts Pavesi - 203 -


projects have been delayed or denied permits because of strong environmental opposition based on
some potential negative environmental effects of these systems.
Keeping the positive aspects in mind, this chapter will review the potential negative
environmental aspects associated with the deployment of a single wind turbine or wind farm. It should
be noted that it ought to be one of the objectives of wind energy system engineers to maximize the
positive impacts while, at the same time, to minimize the negative aspects. In addition to their
importance in the sitting process for wind systems, the potentially adverse environmental effects of
wind turbines tie in directly with much of the wind turbine design material discussed in previous
chapters. With appropriate design of the wind turbine and the wind energy system, these adverse
environmental factors can be minimized.
The potential negative impacts of wind energy can be divided into the following categories:
 Avian interaction with wind turbines
 Visual impact of wind turbines
 Wind turbine noise
 Electromagnetic interference effects of wind turbines
 Land-use impact of wind power systems
 Other impact considerations
At the present time, the first three topics include most of the major environmental issues affecting
wind system deployment, but the other topics are also important and should be addressed. In the next
sections, these potential impacts will be reviewed, including the following general topic areas of each
environmental impact:
 Problem definition
 Source of the problem
 Quantification or measurement of the problem
 Environmental assessment examples
 Reference resources or tools that address the problem
Many of these problems have to be addressed in the permitting phases of a wind project.
Furthermore, depending on the particular site’s environmental assessment regulations, one or
more of these areas may require detailed investigation and/or an environmental impact statement. The
scope of this chapter does not permit a full summary of the various regulations and laws on this subject,
but, where possible, the reader will be referred to the appropriate literature on the subject.

11.1 Life cycle analysis


The wind farm could mitigate an impressive amount of atmospheric emissions. A few questions
that may arise at this point are:
 Don’t we require energy for manufacturing the turbine and constructing the plant?
 Shouldn’t we account the emissions in the manufacturing and commissioning stages of the
system in our analysis?
 Shouldn’t we consider the energy required for disposing the plant after its life period?
 If we consider these factors, can we say that the process of generating energy from wind is
totally free from atmospheric emissions?
 How rapidly could the system recover all the energy consumed for its manufacturing,
installation, operation and dismantling?
To answer these questions, we should broaden our analysis to the entire life period of the project.

| Environmental Aspects and Impacts Pavesi - 204 -


Figure 11.1.1 Various stages of the life cycle of a wind turbine
This is called the life cycle analysis (LSA).
In the life cycle based analysis, we look at a system or technology in its totality and account the
energy use and related emissions involved in all the stages of its production, use and disposal. This
essentially should include the extraction of raw materials, its conversion into different components,
manufacturing, commissioning and operation of the system, and finally its disposal or recycling after
use. For example, the life cycle of a typical wind turbine is shown in Figure 11.1.1. When we assess
such systems, it is logical to account the energy flow and emission potential of all the phases of the
project life as indicated in the figure. Here, upto the phase of turbine manufacturing, energy of various
forms are consumed and thus the system will have a negative impact on environment. In contrast,
during the operational phase, energy is being generated without any pollution and thus the project
reacts positively to the environment.
As all the processes involved in the project-right from the inception to the decommissioning are
considered, the LSA is often termed as a ‘cradle to grave’ method. LSA is a versatile approach for
assessing and comparing the environmental merits of different energy sources and technologies.
Further it can be used as an effective tool for identifying possible improvements in the system design
and decision making.
The LSA broadly consists of five stages as shown in Figure 11.1.1. In the first stage, the goal or
the scope of the analysis is defined. The unit processes involved in the life cycle of the technology
(production, use and disposal) are identified at this stage. The next step is to clearly lay down the
assumptions which form the basis of our analysis. Further, in the inventory analysis stage, the inflow
and outflow of relevant materials and energy, throughout the life cycle of the system, is quantified on
the basis of reliable data. In the impact assessment phase, the environmental implications of the

| Environmental Aspects and Impacts Pavesi - 205 -


material and energy flow Table 13. 3 Materials required for a 600 kW land based wind
are analyzed. For easy turbine
comparison, impacts of Materials kg
different kinds are often Steel 63460
weighed and normalized Aluminum 1780
on the basis of their Copper 740
intensity of environmental Sand 2520
loading. Finally, during Glass 1320
the interpretation and Polyester and epoxy 2400
design improvement Oil products 120
phase, results of the Reinforced iron (foundation) 14400
previous stages are
Concrete (foundation) 339000
analyzed and possible
Miscellaneous 840
design improvements in
the system are identified
Table 13. 4 Additional materials required for the 600 kW offshore
with a detailed plan of
turbine
action.
The first step in the Materials kg
LSA of a wind turbine is Reinforced iron 14400
to quantify the material Concrete 339000
required for the system. Copper 3096
This should include the Lead 4032
materials used for the Steel 4680
fabrication of different PEX 648
turbine components and the foundation. Type and
quantity of materials required for a system varies with
its size (kW rating), design features and site
conditions.
Approximate amount of materials required for a
typical 600 kW land based turbine is given in Table
13. 3. As offshore turbines require stronger
foundation and sea cabling, additional materials are
required for such systems. Materials additional
required for a 600 kW offshore turbine are listed in
Table 13. 4.
Reliability of the LSA greatly depends on the
accuracy in estimating the materials used for the
system. Estimate for turbines of specific size and
design can be obtained from the manufacturers on
request.
There are two common ways to identify the
environmental merit of an energy system under LSA.
They are (1) Net Energy Analysis (NEA) and (2) Life
Cycle Emission (LCE) analysis.

Figure 11.1.2 Various stages involved in LSA

| Environmental Aspects and Impacts Pavesi - 206 -


11.1.1 Net energy analysis
In the Net Energy Analysis, we compare the useful energy produced by the system (E P) with the
energy consumed by it throughout its life cycle (ECL). The ratio between the energy developed and
consumed by the system is often termed as the Energy Payback Ratio (EPR). Thus,
Ep EA L
EPR   (10.7.36)
ECL ECL
where EA is the annual energy production from the system and L is the life period.
All the energy consumed during the manufacturing, operation and disposal phases of the
technology should be included in ECL. Another point of interest is to estimate the time required for the
system to pay back all the energy consumed by it.
This is called the Energy Payback Period (EPP). Thus we have
ECL
EPP  (10.7.37)
EA
The energy produced by the turbine at a given location is basically a function of its capacity
factor. This means that, apart from the turbine type and size, the site characteristics also influence EPR
and EPP. Even with a moderate capacity factor, wind farm projects generally pay back the energy
consumed throughout its life cycle within one year of its commissioning. Modern wind turbines are
more efficient and use fewer materials and hence pay back the energy much quicker than the earlier
designs.
Table 13. 5 Energy equivalent of different
For EPR and EPP calculations, we require
materials used in a 600 kW land based turbine
the amount of energy spent in manufacturing
Materials Energy, MJ
different components of the turbine. Once the
Steel 1630922.0 quantity of different materials required for the
Aluminum 69776.0 components are identified (as in Table 13. 3 and
Copper 57868.0 Table 13. 4), then the energy input in respect of
Sand 9273.6 these components can be calculated by
Glass 12276.0 multiplying the weight of the materials used by the
Polyester and epoxy 109680.0 specific energy required to produce them. The
Reinforced iron (foundation) 522720.0 specific energy depends on the energy consumed
Concrete (foundation) 1247520.0 from various sources in producing these materials.
Miscellaneous 39228.0 For example, if the energy consumption in
Totale 3699263.6 producing one kg of steel is 1.6 MJ equivalent of
coke, 17.4 MJ equivalent of coal, 6.6 MJ
Table 13. 6 Energy requirement of additional equivalent of oil and 0.1 MJ equivalent of natural
materials used in a 600 kW offshore turbine gas, then the specific energy of steel is 25.7
Materials Energy, MJ MJ/kg.
Reinforced iron 522720.0 Energy use pattern of a 600 kW wind turbine
Concrete 1247520.0 is given in Table 13. 5 and Table 13. 6. This
Copper 242107.2 includes the energy associated with the
Lead 143539.2 production, transportation and manufacture of the
Steel 120276.0 materials.
Totale 2276162.4 The energy required to produce the materials

| Environmental Aspects and Impacts Pavesi - 207 -


may differ from place to place, depending on the locally available energy source. When we have a
different situation-for example, in a country where hydro or nuclear power are the major sources-the
scenario will change accordingly.
The technology adopted for producing the materials is another factor influencing the energy use.
For example, energy equivalent of steel can vary from 20.7 MJ/kg to 30.6 MJ/kg, based on the nature
of the processes involved in its production. Thus, it should be ensured that the values assumed in our
calculations are valid for the specific region of interest.
From the above tables, it is seen that the total energy consumed by a land based wind farm, in
respect of its material of construction is 3,699.6 GJ. For an offshore system, the corresponding energy
requirement is 5,975.5 GJ. Wind turbines do not consume energy during the operational phase of its
life cycle. However, energy is required for decommissioning the system after its life period as we
require energy for dismantling, transporting, and disposal or recycling the components. The energy
demand in this regard can be taken as 0.1-0.4 MJ/kg [Schleisner L (2000)] of the materials used. Thus,
the energy required for disposing the 600 kW land based turbine is 106.6 GJ where as that for the
offshore units is 198.1 GJ. This makes the total energy requirement of the land based turbine, during its
life cycle as 3806.2 GJ. For the offshore system, corresponding energy requirement is 6173.6 GJ.
Hence, the wind farm would pay back all the energy consumed in its life cycle within 8 months
of its commissioning.
The Energy Payback Ratio and Energy Payback Period of different sources are compared in
andFigure 11.1.3. Here, the life period of the coal, fusion and fission projects are assumed to be 40
years and that for gas and PV technologies are taken as 30 years. For coal, gas, fusion and fission
technologies, capacity factor of 75 per cent is assumed. Similarly, the conversion efficiency of PV
modules are taken as 6 per cent. Comparing various technologies, we can see that wind energy has the
highest EPR and lowest EPP, which is a clear indication of the environmental merit of wind energy
over other options.

| Environmental Aspects and Impacts Pavesi - 208 -


Figure 11.1.4 Energy payback period of various energy sources
The offshore project has lower EPR and higher EPP in comparison with the land based system.
These differences are due to the extra materials used for the foundation of the offshore project.

Figure 11.1.3 Energy Payback Ratio of various energy sources.

| Environmental Aspects and Impacts Pavesi - 209 -


11.1.2 Life cycle emission
Wind energy do not pose the threat of atmospheric emission during its energy generation phase.
However, energy in various forms is being consumed during the construction, commissioning and
decommissioning phases of the project. In the Life Cycle Emission (LSE) analysis, we account
emissions during all these phases of the turbine’s life cycle.
Possible emissions corresponding to the materials used for 600 kW turbine, assuming the Danish
conditions, are given in Error! Reference source not found., Error! Reference source not found.
and Error! Reference source not found..
We can see that the major pollutants here are CO2, SO2 and NOx. Emission of N2O, CH4, CO and
NMVOC are marginal. Life cycle emissions of wind energy, as compared with other sources are shown
in Table 11.1.2. Even if we consider all possible emissions throughout its life cycle, wind emerges out
as the cleanest source of energy available today.

Table 11.1.2 Emissions from additional materials required for the offshore wind
turbine
Materials CO NMVOC CH4 N2O CO2 NOX SO2
Reinforced iron 22.61 2.59 0.86 1.3 44841.6 128.02 209.95
Concrete 0 0 0 0 238317 847.5 3.39
Copper 4.86 0.77 0.5 0.59 20235.46 71.8 110.25
Lead 8.18 2.22 0.28 0.44 11906.5 79.91 73.34
Steel 4.35 0.75 0.19 0.33 10794.42 44.46 67.86
Total 40.01 6.33 1.83 2.66 326094.97 1171.69 464.79

Table 11.1.3 Emissions due to the disposal of materials


Materials SO2 NOX CO2 N2O CH4 NMVOC CO
Polyester and epoxy 8.64 14.4 11232 0.38 0.58 0.86 210.05
Miscellaneous 3.02 5.04 3931.2 0.13 0.2 0.3 73.52

Table 11.1.1 Emissions from different materials used for a 600 kW land based wind
turbine.
Materials SO2 NOX CO2 N2O CH4 NMVOC CO
Steel 920.17 602.87 146370.49 4.44 2.54 10.15 59.02
Aluminum 37.38 23.14 6111.63 0.19 0.12 0.26 1.33
Copper 26.35 17.16 4836.64 0.14 0.12 0.19 1.16
Glass 1.15 3.18 766.92 0.01 0.05 0.2 0.87
Polyester and epoxy 54.98 35.3 9458.4 0.29 0.19 0.48 2.64
Reinforced iron 209.95 128.02 44841.6 1.3 0.86 2.59 22.61
Concrete 3.39 847.5 238317 0 0 0 0
Total 1253.38 1657.17 450702.68 6.37 3.88 13.87 87.62

| Environmental Aspects and Impacts Pavesi - 210 -


Figure 11.1.5 Comparison of life cycle emissions from various energy sources

| Environmental Aspects and Impacts Pavesi - 211 -


11.2 Avian Interaction with Wind Turbines
11.2.1 Overview of the problem
The environmental problems associated with avian interaction and wind systems surfaced in the
United States in the late 1980s. For example, it was found that birds, especially federally protected
golden eagles and red-tailed hawks, were being killed by wind turbines and high voltage transmission
lines in wind farms in California’s Altamont pass. This information caused opposition to the Altamont
Pass project among many environmental activists and aroused the concern of the US Fish and Wildlife
Service, which is responsible for enforcing federal species protection laws.
There are two primary concerns related to this environmental issue: (1) effects on bird
populations from the deaths caused by wind turbines, and (2) violations of the Migratory Bird Treaty
Act, and/or the Endangered Species Act. These concerns, however, are not confined to the United
States alone. Bird kill problems have surfaced at other locations in the world. For example, in Europe,
major bird kills have been reported in Tarifa, Spain (a major point for bird migration across the
Mediterranean Sea), and at some wind plants in Northern Europe. Wind energy development can
adversely affect birds in the following adverse manners (Colson, 1995):
 Bird electrocution and collision mortality
 Change to bird foraging habits
 Alteration of migration habits
 Reduction of available habitat
 Disturbance of breeding, nesting, and foraging
It should also be pointed out that the same author states that wind energy development has the
following beneficial effects on birds:
 Protection of land from more dramatic habitat loss
 Provision of perch sites for roosting and hunting
 Provision and protection of nest sites on towers and ancillary facilities
 Protection or expansion of prey base
 Protection of birds from indiscriminate harassment
At the present time, the long-term implications of the bird issue on the wind industry are not
clear. Problems may arise in areas where large numbers of birds congregate or migrate, as in Tarifa, or
where endangered species are affected, as in Altamont Pass (NWCC, 1998).
This could include a large number of locations, however, since some of the traits that characterize
a good wind site also happen to be attractive to birds. For example, mountain passes are frequently
windy because they provide a wind channel through a mountain range.
For the same reason, many times they are the preferred routes for migratory birds.

11.2.2 Characterization of the problem


There is a close correlation between a locality and its bird fauna. Many bird species are very
habitat-specific and often particularly sensitive to habitat changes. Furthermore, there is a
correspondingly close correlation between the site and the location of the wind turbines, which is
especially dependent on the wind conditions. In European (Clausager and Nohr, 1996) and US studies
(Orloff and Rannery, 1992) the impact on birds is generally divided into the following two categories:
(1) direct impacts, including risk of collision, and (2) indirect impacts, including other disturbances

| Environmental Aspects and Impacts Pavesi - 212 -


from wind turbines (e.g., noise). Such effects may result in a total or partial displacement of the birds
from their habitats and deterioration or destruction of the habitats.
The risk of collision is the most obvious direct effect, and numerous studies have focused on it.
These studies include the estimation of bird numbers that collide with the rotor or associated structures,
and they have concentrated on the development of methods for the analysis of the extent of the
collisions. Indirect effects include:
 Disturbance of breeding birds
 Disturbance of staging and foraging birds
 Disturbing impact on migrating and flying birds

11.3 Current mitigation concepts and case studies


11.3.1 Mitigation concepts
A number of recent studies have specifically addressed the subject of measures for the
minimization of the impact of wind turbine systems on birds. The most detailed work has taken place in
California and has been supported by the California Energy Commission and the American Wind
Energy Association. Typical mitigation measures from some of these studies (Colson, 1995; and Wolf,
1995) include the following:
 Avoid migration corridors. Major bird migration corridors and areas of high bird
concentrations should generally be avoided when sitting wind facilities unless there is
evidence that because of local flight patterns, low bird usage of the specific area, or other
factors, the risk of bird mortality is low
 Fewer, larger turbines. For a desired energy capacity, fewer large turbines may be
preferred over many smaller turbines to reduce the number of structures in the wind farm
Avoid micro habitats. Microhabitats or fly zones should be avoided when sitting
individual wind turbines.
 Alternate tower designs. Where feasible, tower designs that offer few or no perch sites
should be employed in preference to lattice towers with horizontal members that are Wind
suitable for perching. Unguyed meteorological and electrical distribution poles should be
used over guyed structures. Also, existing lattice towers should be modified to reduce
perching opportunities
 Remove nests. With agency approvals, raptor nests found on structures should be moved
to suitable habitat away from wind energy facilities Prey base management. Where
appropriate, prey base management should be investigated as an option. A humane
program, such as live trappings should be established to remove unwanted prey from
existing wind farms
 Bury electrical lines. Electrical utility lines should be underground when feasible. New
overhead electrical distribution systems should be designed to prevent bird electrocutions.
Existing facilities should employ techniques that significantly reduce bird electrocutions
 Site specific mitigation studies. The cause and effect of bird interactions with wind energy
facilities should be examined to determine the manner in which collisions occur and
appropriate mitigation identified and employed
 Conservation of alternative habitats. Habitat conditions that will benefit affected species
should be maintained to protect them from other more damaging land-uses

| Environmental Aspects and Impacts Pavesi - 213 -


11.3.2 Case studies
Numerous technical reports and papers exist that have focused on the potential or actual
environmental effects of wind turbine systems on birds at a specific site. Although it is beyond the
scope of this text to review them in detail, a few of the more appropriate ones (and their particular
applications) are listed below:
 California wind farm studies. A review of this subject, as applied to the wind farms in
Altamont Pass and Solano County, is contained in a technical report by the California
Energy Commission (Orloff and Hannery, 1992).
 Recent US studies. Sinclair (1999) and Sinclair and Morrison (197) give an overview of
the recent US avian research program through 1999. Also, Goodman (1997) presents a
summary of the avian studies (and other environmental issues) carried out for a 6 MW
wind farm in Vermont.
 European studies. Gipe (1995) presents an overview of bird-related environmental impact
studies in Europe up to 1995. In addition, representative examples of European
environment impact studies of birds (especially for Denmark and the U.K.) are given in
the technical papers of Clausager and Nohr (1996) and Still et al. (1996).

11.4 Resources for environmental assessment studies of avian impacts


A review of the previously referenced case studies regarding the environmental impacts of wind
systems on birds reveals that such studies can require specialized expertise, be very detailed, and can
add to the cost and deployment time of a potential wind site. As mentioned previously, a representative
example of this process is given in the technical report of the California Energy Commission (Orloff
and Flannery, 1992). In general, this type of work can be divided into two parts: (1) a complete
definition of the study area, and (2) an assessment of bird risk.
Once the site for a wind plant is identified, the potential effect of the proposed wind farm
development on the birds of concern should be systematically studied.
If the available information in this regard is not sufficient to establish the biological suitability of
the site, we should conduct onsite monitoring and surveys using appropriate methods and metrics. All
the bird groups of special concern, including the breeding, migrating and wintering species, should be
included in such studies.
Apart from the mere number of birds and its distribution in the wind farm area, the functional
behaviour of the species in the region should also be included in the analysis.
The most common experimental design adopted in such investigations is the Before-After Control
Impact (BACI) [Howell JA (1997)]. Under this method, data on avian activities are collected from the
wind site before and after the construction of the farm as well as from a controlled reference site. The
activities are recorded using systematically planned point count surveys. The bird fatalities are
estimated through carcass searches around some systematically or randomly selected turbine areas.
Although the true population size is not included in such studies, trends and indices that can be
correlated with actual population parameters can be derived.
National Wind Coordination Committee (NWCC) has prepared the procedures for conducting
such investigations [Anderson R et al. (1999)]. Results of such studies can be used to develop an index
for risk of collision of birds with wind turbines. The collision risk index for a given bird species is
given by [Erickson WP et al. (2000)]
IR  U Pf Pt (10.7.38)

| Environmental Aspects and Impacts Pavesi - 214 -


Where U is the mean use by a species adjusted for visibility bias, P f is the fraction of the total
observations in which the activity of the subject is ‘flying’, and Pt is the fraction of all the flight height
observations coming under the height band swept by the turbines.
It should be ensured that the risk index is under the limit permitted by the environmental laws
framed by the regulatory bodies.
The avian risk at a given site can be reduced to some extent by the optimum layout of the wind
farm and configuration of individual turbines. If the migration route of the birds and the corresponding
climatic conditions are identified, the turbines can be laid out in such a way that clear spaces are
available for the birds to pass through. In areas of migration, the turbines can be widely spaced leaving
clear route for the migration and thus the risk of bird collision can be minimized.
Visual acuity of birds is a critical factor influencing the avian-wind turbine interaction.
Considerable research in this aspect has been done by Dr. Hugh McIsaac [Morrison ML (2000)].
He found that the raptors can see the turbine blades having an average width of 0.6 m, even from a
distance of 1000 m. However, as the blades start rotating, the visibility reduces as a result of blurring.
Contrast is another factor influencing the visibility. In order to make the moving components distinctly
visible, the contrast between the blades and the background should be high. Thus to make the turbine
blades maximally visible, the turbine blades may be painted in a highly contrasting pattern.
Noise can also be used as an effective tool for making the turbine more detectable for birds,
especially during the night hours. Modern turbines are less noisy and the inherent noise level of these
turbines does not keep the birds away. Efforts are even made to attach separate devices to the rotor tips
which can generate noise in levels that can startle birds and draw their attention. Different types of bird
flight diverters like balls and spirals as used in power lines are also being experimented.
Horizontal axis turbines mounted on horizontal lattice towers are more susceptible to bird
collision. Further, these types of tower structures with flat crossbars provide ideal platform for perching
birds. Tubular structures are less preferred by the birds for perching and hence could reduce the
probability of wind turbine-bird interaction. Most of the modern wind turbines are provided with
tubular structures.
Collision of birds with wind turbines is not as severe an issue as it is often being projected. It is
reported that a bird in the vicinity of a farm will normally collide with the turbines for not more than
once in every 8 to 15 years [Benner JBH (1992)]. As a matter of fact, birds may collide with any man
made structures. For example, around 1.25 million birds die every year due to collision with tall
structures like buildings and towers. Similarly, 97.5 million birds are killed as they strike with glass.
Estimates of bird mortality due to hitting vehicles are more alarming – 57 million a year [Kenetech
Wind power (1994)]. In comparison with these causalities, wind turbines have only a marginal impact
on avian population. Accessory structures and distribution networks often pose more threat to birds
than the wind turbines itself. In a study spanned for over four years [Hunt WG (2000)], it was found
that fifty four per cent of the mortalities were due to the electric generation and transmission structures.
Causalities due to the wind turbine blade striking were only 38 per cent.
While discussing the avian issues of wind turbines, it is logical to consider the effects of other
generation technologies on the environment. For example, the green house emissions, acid rain and
smog resulting from the coal based plants have more far reaching effects on the environment and the
ecosystem, than the wind turbines. This already destroyed many lakes, streams and forests all around
the globe. Changes in the weather pattern due to the global warming threaten many climate sensitive
species. Apart from the risk from radioactive wastes, the hot water discharged from the nuclear plants
has adverse effect on the marine habitat. For example, San Onofre nuclear station in California is

| Environmental Aspects and Impacts Pavesi - 215 -


responsible for killing around 21 tons of fish every year, including several billions of fish larvae [NIRS
(1989)]. A single accident from these plants may wipe out thousands and thousands of birds and other
protected species from the earth. Structures of the conventional generating plants (towers, electric lines
etc.) also pose threat to birds and bats, which are often left unaccounted.
Although relatively small numbers of birds are killed due to collision with wind turbines, the
wind energy communities all around the world are making every effort to resolve this issue of avian
mortality. For example, the American Wind Energy Association (AWEA), in coordination with the
research organizations and industrial groups, has established an environmental task force to tackle this
problem.

| Environmental Aspects and Impacts Pavesi - 216 -


Similarly, in UK, the Energy Technology Support Unit (ETSU) under the Department of Trade

Figure 11.4.1 Eight categories of wind turbines in the Altamont pass (Orloff and Flannery,
1992)

| Environmental Aspects and Impacts Pavesi - 217 -


and Industry has initiated activities to reduce the avian deaths due to wind energy projects. Findings of
several studies made in this direction is compiled in [Lowther SM, Tyler S (1996)].
The first part primarily consists of a detailed definition of the site topography and machine sitting
layout of the wind farm. In addition, when different types of wind turbines and towers are used, it is
important to group turbines into different categories. For example, Figure 11.4.1. illustrates the eight
categories of wind turbines that were installed at the Altamont Pass in California in the 1980s. Another
part of this phase of the work involves the selection of sample sites where detailed bird - wind turbine
interaction data are to be collected.
The second, and much more detailed part of this type of study, the assessment of bird risk,
generally involves a comprehensive methodology for accomplishing this task. In the United States, a
number of sponsors including the National Renewable Energy Laboratory (NREL) and the National
Wind Coordinating Committee (NWCC) Avian subcommittee have recently worked to develop this
type of methodology. This work has focused on the development of a standardized method for
determination of the factors responsible for avian deaths from wind turbine facilities, and scientific
methods that can be used to reduce fatalities. The following summary presents some of the important
methods, measurements, and relationships (metrics) that have been used for this type of work
(Anderson et al., 1997; NWCC, 1999):
Bird utilization counts. Under this part of a study, an observer notes the location, behaviour, and
number of birds using an area. This is done in repeatable ways, using standard methods, so that results
can be compared with bird utilization counts from other studies.
Bird behaviours to be noted include: flying, perching, soaring, hunting, foraging, and actions
close (50 meters or less) to wind farm structures.
Bird utilization rate. The bird utilization rate is defined as the number of birds using the area
during a given time. It is based on bird utilization count. Thus:
n of Birds Observed
Bird Utilization Rate  (10.7.39)
Time
or
n of Birds Observed
Bird Utilization Rate  (10.7.40)
Time  Area
Bird mortality. The bird mortality is defined as the number of observed deaths, per unit area
(again, this is based on the bird utilization count). Therefore:
n of Dead Birds
Bird Mortality  (10.7.41)
Defined Search Area
Bird risk. Bird risk is a measure of the likelihood that a bird using the area in question will be
killed. It is defined as follows:
Bird Mortality n of Dead Birds Defined Area
Bird Risk   (10.7.42)
Bird Utilization Rate Defined Search Area Time
Bird Risk can be used to compare risk differences for many different variables: i.e., distances
from wind facilities; species, type, and all birds observed; seasons; and turbine structure types. It can be
used to compare risks between wind resource areas and with other types of facilities such as highways,

| Environmental Aspects and Impacts Pavesi - 218 -


power lines, and TV and radio transmission towers.
Rotor swept area metrics. Under this category two measurements that consider the effects of
different wind turbine sizes and designs have been defined. The first, rotor swept hours is defined as:
Rotor Swept Hours  Rotor Swept Area  Hours of Operation (10.7.43)
This parameter combines the size of the area of the rotor with the time it operates. The second
parameter, rotor swept hour risk (RSHR) allows for a comparison of risks associated with different
rotor swept areas, or turbine sizes, in relation to the time they operate, It is defined by:
Bird Risk
Rotor Swept Hour Risk  (10.7.44)
Rotor Swept Hours
A detailed discussion of the example use of these metrics is beyond the scope of this work. It is
expected that many of them will be used in future research studies aimed at developing methods to
measure the risk to birds not only from wind systems, but also from other human created hazards such
as buildings and highways (NWCC, 1998).
In summary, it should be pointed out that even if the initial research indicates that a wind energy
project is unlikely to seriously affect bird populations, further studies might be needed to verify this
conclusion. These could include monitoring baseline bird populations and behaviour before the project
begins, then simultaneously observing both a control area and the wind site during construction and
initial operation. In certain cases, operational monitoring might have to continue for years.

11.5 Visual Impact of Wind Turbines


One of wind power’s perceived adverse environmental impact factors, and a major concern of the
public, is its visibility (Gipe, 1995). Compared to the other environmental impacts associated with wind
power, the visual impact is the least quantifiable. For example, the public’s perceptions may change
with knowledge of the technology, location of wind turbines, and many other factors. Although the
assessment of a landscape is somewhat subjective, professionals working in this area are trained to
make judgments on visual impact based on their knowledge of the properties of visual composition and
by identifying elements such as visual clarity, harmony, balance, focus, order, and hierarchy (Stanton,
1995).
Wind turbines need to be sited in well exposed sites in order to be cost effective. It is also
important for a wind engineer to realize that the visual appearance of a wind turbine or a wind farm
must be considered in the design process at an early stage. For example, the degree of visual impact is
influenced by such factors as the type of landscape, the number and design of turbines, the pattern of
their arrangement, their colour, and the number of blades.

11.5.1 Characterization of the problem


Visual or aesthetic resources refer to the natural and cultural features of an environmental setting
that are of visual interest to the public. An assessment of a wind project’s visual compatibility with the
character of the project setting can be based on a comparison of the setting and surrounding features
with simulated views of the proposed project. in this light, the following parameters and questions can
be considered (NWCC, 1998):
 View shed alteration: will the project substantially alter the existing project setting
(generally referred to as the ‘view shed’), including any changes in the natural terrain or

| Environmental Aspects and Impacts Pavesi - 219 -


landscape?
 View shed consistency: will the project deviate substantially from the form, line, colour,
and texture of existing elements of the view shed that contribute to its visual quality?
 View shed degradation: will the project substantially degrade the visual quality of the
view shed, affect the use or visual experience of the area, or intrude upon or block views
of valuable visual resources?
 Conflict with public preference: will the project be in conflict with directly identified
public preferences regarding visual and environmental resources?
 Guideline compatibility: will the project comply with local goals, policies, designations,
or guidelines related to visual quality?
The scope of a visual impact assessment depends on whether a single or a number of wind
turbines are to be sited in a particular location. Visual impact is not directly proportional to the number
of turbines in a wind farm development. It will, however, vary a great deal between a single turbine and
a wind farm. That is, a single wind turbine has only a visual relationship between itself and the
landscape, but a wind farm has a visual relationship between each turbine as well as with the landscape
(Stanton, 1995).

| Environmental Aspects and Impacts Pavesi - 220 -


Harmony
& Clarity

Order

A Point

A Line &
Edge

Parallel
Lines

Grid

Clusters
Groups
Rhythm
&
Repetition

Shape

Figure 11.5.1 Summary of fundamental design principles significant to wind farm development
(Stanton, 1994).
An important task here is the characterization of the location where the proposed wind system is
to be sited. As one example of this process, the work of Stanton (1994) as applied to wind farms in the
U.K uses the design principles shown in Figure 11.5.1. It should be noted that although this visual
impact study is based on widely known and accepted principles (such as those summarized in Figure
11.5.1), the judgements made are subjective in nature.

11.5.2 Design of wind systems to minimize visual impact


In the United States and Europe, there are numerous publications that suggest potential designs
for wind systems that minimize visual impact (e.g., see Ratto and Solari, 1998). In many cases, the
subjective nature of this topic appears and there are major differences of opinion between researchers.
An example of this problem includes the choice of turbine colours (Aubrey, 2000). Two examples of
previous design strategies (one from the United States and the other European) for the reduction of the
visual impact of wind energy systems follow.

11.5.3 Visual impact mitigation design strategies for US sites

| Environmental Aspects and Impacts Pavesi - 221 -


From work carried out in the US (NWCC, 1998), a number of design strategies for reducing the
environmental impact of wind systems have been proposed. They include the following:
 Using the local land form to minimize visibility of access and service roads, and to protect
land from erosion
 Use of low profile and unobtrusive building designs to minimize the urbanized appearance
or industrial character of projects located in rural or remote regions
 Use of uniform colour, structure types, and surface finishes to minimize project visibility
in sensitive areas with high open spaces. Note, however, that the use of non-obtrusive
designs and colours may conflict with efforts to reduce avian collisions and may be in
direct conflict with aircraft safety requirements for distinctive markings
 Selecting the route and type of support structures for above-ground electrical facilities as
well as the method, mode, and type of installation (below- vs. above-ground). Where
multiple generation units are to be sited close together, consolidation of electrical lines
and roads into a single right of way, trench, or corridor will cause fewer impacts than
providing separate access to each unit
 Controlling the placement and limiting the size, colour, and number of label markings
placed on individual turbines or advertising signs on fences and facilities
 Prohibiting lighting, except where required for aircraft safety, prevents light pollution in
otherwise dark settings. This may incidentally minimize collision by nocturnal feeders
which prey on insects attracted to lights
 Controlling the relative location of different turbine types, densities, and layout geometry
to minimize visual impacts and conflicts. Different turbine types and those with opposing
rotation can be segregated by buffer zones. Mixing of types should be avoided or
minimized

11.5.4 European wind farm visual impact design characteristics


The assessment of the visual impact of a proposed wind farm development is an important design
step in recent European work. One example of this type of work is summarized by the following list of
wind farm design characteristics quoted from Stanton’s (1994, 1995) U.K. studies:
 Wind turbine form
 Blade number
 Turbine nacelle and tower
 Turbine size
 Wind farm size
 Spacing and layout of wind turbines
 Colour
Note that this work complements Stanton’s previously described landscape character types.
Again, the subjective nature of this example should be emphasized.

11.5.5 Resources for visual impact studies


In both the US and Europe there are numerous references and handbooks where one can find
design guidelines for minimizing the visual environmental impact of wind turbines and wind farms
(e.g., see Ratto and Solari, 1998). A summary of some of the most representative examples follows.

| Environmental Aspects and Impacts Pavesi - 222 -


11.5.6 US visual impact resources
A good resource on the subject is contained in a handbook for the permitting of wind energy
facilities (NWCC, 1998). Some of the guidelines developed in this handbook have been previously
discussed. In addition, this reference contains much additional information of potential use in this and
many other phases of environmental impact assessment. For example, the authors point out that a
valuable process tool for the assessment of potential project impacts to sensitive visual resources is the
preparation and use of visual simulations.
Goodman (1997) gives another good source of information on visual resources. This paper
documents the environmental impacts of a wind farm installation in rural Vermont, and summarizes the
methods used to minimize the environmental impact of a 6 MW wind farm.

11.5.7 European visual impact resources


As discussed in numerous publications on the subject, the visual impact of wind turbines and
wind farms have been under significant study in many European countries. A great deal of this work
has been carried out in the U.K. For example, the visual impact assessment work of Stanton
(1994,1995) has already been mentioned.
In Europe, a number of investigators have developed some very sophisticated and useful
techniques that can be used to illustrate the visual intrusion of a potential wind farm installation.
Furthermore, many of these have been successfully used in actual wind farm development projects (see
Ratto and Solari, 1998). For example, one uses digital computer techniques based on topographic
information and wind turbine design characteristics.
With this type of technique it is possible to plot ‘zones of visual impact’ on a map to illustrate the
locations where the development might be seen. One disadvantage to this technique is that it takes no
account of local screening, e.g., by buildings or trees, so it is a worst case scenario. An example of the
use of this type of technique, using geographical information systems (GIS), is given by Kidner (1996).
Another method (see Taylor and Durie, 1995) is based on the use of photomontages. Here, use is
made of panoramic photographs taken from certain key locations of varying distances from the
proposed site, and superimposing suitably scaled wind turbines in the photographs. A limitation to this
technique is that the visibility of a wind system will change with time of day, season, and with certain
weather conditions - especially when the turbine blades are rotating. A third method attempts to
overcome some of these disadvantages by creating a video montage of a proposed site and then super-
imposing rotating wind turbines on the scene (Robotham, 1992). While this creates a most effective
visual presentation, its major disadvantage is its high cost compared to the other two methods. Today, a
number of commercial software packages now include all or some of these capabilities.

11.6 Wind Turbine Noise


The problems associated with wind turbine noise have certainly been one of the more studied
environmental impact areas in wind energy engineering. Noise levels can be measured, but, as with
other environmental concerns, the public’s perception of the noise impact of wind turbines is a partly
subjective determination.
Noise is defined as any unwanted sound, Concerns about noise depend on the level of intensity,
frequency, frequency distribution and patterns of the noise source; background noise levels; terrain
between emitter and receptor; and the nature of the noise receptor. The effects of noise on people is
classified into three general categories (WNCC, 1998):

| Environmental Aspects and Impacts Pavesi - 223 -


 Subjective effects including annoyance, nuisance, dissatisfaction
 Interference with activities such as speech, sleep, and learning
 Physiological effects such as anxiety, tinnitus (ringing in ear), or hearing loss
In almost all cases, the sound levels associated with environmental noise produce effects only in
the first two categories. Workers in industrial plants and those who work around aircraft can experience
noise effects in the third category. Whether a noise is objectionable will depend on the type of noise
(see the next section) and the circumstances and sensitivity of the person (or receptor) who hears it.
Mainly because of the wide variation in the levels of individual tolerance for noise, there is no
completely satisfactory way to measure the subjective effects of noise, or of the corresponding
reactions of annoyance and dissatisfaction.
Operating noise produced from wind turbines is considerably different in level and nature than
most large-scale power plants, which can be classified as industrial sources.
Wind turbines are often sited in rural or remote areas that have a corresponding ambient noise
character. Furthermore, while noise may be a concern to the public living near wind turbines, much of
the noise emitted from the turbines is masked by ambient or the background noise of the wind itself.
The noise produced by wind turbines has diminished as the technology has improved.
For example, with improvements in blade airfoils and turbine operating strategy, more of the
wind energy is converted into rotational energy, and less into acoustic noise. Even a well designed
turbine, however, can generate some noise from the gearbox, brake, hydraulic components or even
electronic devices.
The significant factors relevant to the potential environmental impact of wind turbine noise are
shown in Figure 10.3 (Hubbard and Shepherd, 1990). Note that this technology is based on the
following primary elements: noise sources, propagation paths, and receivers. In the following sections,
after a short summary of the basic principles of sound and its measurement, a review of noise
generation from wind turbines, its prediction, and propagation, as well as noise reduction methods are
given.

11.6.1 Noise and sound fundamentals


11.6.1.1 Characteristics of sound and noise

Sound is generated by numerous mechanisms and is always associated with rapid small scale
pressure fluctuations (which produce sensations in the human ear). Sound waves are characterized in
terms of their wavelength, A, frequency, f , and velocity U, where U is found from:
U=f (10.9.1)
The velocity of sound is a function of the medium through which it travels, and it generally
travels faster in denser media. The velocity of sound is about 340 m/s in atmospheric air. Sound
frequency determines the ‘note’ or pitch that one hears, which, in many cases, corresponds to notes on
the musical scale (Middle C is 262 Hz).
An octave denotes the frequency range between sound with one frequency and one with twice
that frequency. The human hearing frequency range is quite wide, generally ranging from about 20 Hz
to 20 MHz. Sounds experienced in daily life are usually not a single frequency, but are formed from a
mixture of numerous frequencies from numerous sources.
Sound turns into noise when it is unwanted. Whether sound is perceived as a noise depends on the
response to subjective factors such as the level and duration of the sound.

| Environmental Aspects and Impacts Pavesi - 224 -


There are numerous physical quantities that have been defined which enables sounds to be
compared and classified, and which also give indications for the human perception of sound.
They are discussed in numerous texts on the subject (for wind turbine noise see Wagner et al.,
1996) and are reviewed in the following sections.

11.6.1.2 Sound power and pressure measurement scales

It is important to distinguish between sound power level and sound pressure level. Sound power
level is a property of the source of the sound and it gives the total acoustic power emitted by the source.
Sound pressure level is a property of sound at a given observer location and can be measured there by a
single microphone. In practice, the magnitude of an acoustical quantity is given in logarithmic form,
expressed as a level in decibels (dB) above or below a zero reference level. For example, using
conventional notation, a 0 dB sound power level will yield a 0 dB sound pressure level at a distance of
I m.
Because of the wide range of sound pressures to which the ear responds (a ratio of 105 or more
for a normal person), sound pressure is an inconvenient quantity to use in graphs and tables. In
addition, the human ear does not respond linearly to the amplitude of sound pressure, and, to
approximate it, the scale used to characterize the sound power or pressure amplitude of sound is
logarithmic (see Beranek and Ver, 1992).
The sound power level of a source, L,, in units of decibels (dB), is given by:
Lw =10 log10(W/W0) (10.9.2)
-12
where W is the source sound power and W0 is a reference sound power (usually 10 W ).
The sound pressure level of a noise. Lp , in units of decibels (dB), is given by:
Lp =10 log10(p/p0) (10.9.3)
where p is the instantaneous sound pressure and p0 a reference sound pressure (usually 20 10-5 Pa)

| Environmental Aspects and Impacts Pavesi - 225 -


The threshold of pain for the human ear is about 200 Pa, which corresponds to a sound pressure
level of 140 dB.

11.6.1.3 Measurement of sound or noise

Sound pressure levels are measured via the use of sound level meters. These devices make use of
a microphone, which converts pressure variations to a voltage signal, which is then recorded on a
meter (calibrated in decibels). The decibel scale is logarithmic and has the following characteristics
(NWCC, 1998):
 Except under laboratory conditions, a change in sound level of 1 dB cannot be perceived
 Outside of the laboratory, a 3 dB change in sound level is considered a barely discernible
difference
 A change in sound level of 5 dB will typically result in a noticeable community response.
 A 10 dB increase is subjectively heard as an approximate doubling in loudness, and
almost always causes an adverse community response
A sound level
measurement that
combines all frequencies
into a single weighted
reading is defined as a
broadband sound level.
For the determination of
the human ear’s response
to changes in noise, sound
level meters are generally

Figure 11.6.2 Definition of A, B, and C frequency weighting scales


(Beranek and Ver, 1992)

Figure 11.6.1 Wind turbine noise assessment factors (Hubbard and Shepherd, 1990)

| Environmental Aspects and Impacts Pavesi - 226 -


equipped with filters that give less weight to the lower frequencies. As shown in Figure 11.6.2, there
are a number of filters (referenced to A, B, and C) that accomplish this. The most common scale used
for environmental noise assessment is the A scale. Measurements made using this filter are expressed
in units of dB(A). Beranek and Ver (1992) discuss details of these scales.
Once the A weighted sound pressure is measured over a period of time, it is possible to determine
a number of statistical descriptions of time-varying sound and to account for the greater sensitivity to
nigh time noise levels. Common descriptors include:
 L10, L50, L90. The A-weighted noise levels that are exceeded 10%, 50%, and 90% of the
time, respectively. During the measurement period L90 is generally taken as the
background noise level.
 Leq (equivalent noise level). The average A-weighted sound pressure level which gives the
same total energy as the varying sound level during the measurement period of time.
 Ldn (day-night noise level). The average A-weighted noise level during a 24 hour day,
obtained after addition of 10 dB to levels measured in the night between 10 p.m. and 7
a.m.

11.6.1.4 Noise mechanisms of wind turbines

There are four types of noise that can be generated by wind turbine operation: tonal, broadband,
low-frequency, and impulsive. They are described below:
 Tonal. Tonal noise is defined as noise at discrete frequencies. It is caused by wind turbine
components such as meshing gears, non-linear boundary layer instabilities interacting with
a rotor blade surface, by vortex shedding from a blunt trailing edge, or unstable flows over
holes or slits
 Broadband. This is noise characterized by a continuous distribution of sound pressure with
frequencies greater than 100 Hz. It is often caused by the interaction of wind turbine
blades with atmospheric turbulence, and is also described as a characteristic ‘swishing’ or
‘whooshing’ sound
 Low frequency. This describes noise with frequencies in the range of 20 to 100 Hz mostly
associated with downwind turbines. It is caused when the turbine blade encounters
localized flow deficiencies due to the flow around a tower, wakes shed from other blades,
etc
 Impulsive. Short acoustic impulses or thumping sounds that vary in amplitude with time
characterize this noise. They may be caused by the interaction of wind turbine blades with
disturbed air flow around the tower of a downwind machine, and/or the sudden
deployment of tip breaks or actuators
The cause(s) of noise emitted from operating wind turbines can be divided into two categories:
(1) aerodynamic and (2) mechanical. Aerodynamic noise is produced by the flow of air over the blades.
The primary sources of mechanical noise are the gearbox and the generator. Mechanical noise is
transmitted along the structure of the turbine and is radiated from its surfaces. A summary of each of
these noise mechanisms follows. A more detailed review is included in the text of Wagner et al. (1996).

| Environmental Aspects and Impacts Pavesi - 227 -


11.6.1.5 Aerodynamic noise

Aerodynamic noise originates from the flow of air around the blades. As shown in Figure
11.6.3(Wagner et al., 1996), a large number of complex flow phenomena generate this type of noise.
This type of noise generally increases with tip speed or tip speed ratio. It is broadband in character and
is typically the largest source of wind turbine noise. When the wind is turbulent, the blades can emit
low-frequency noise as they are buffeted by changing winds. If the wind is disturbed by flow around or
through a tower before hitting the blades (on a downwind turbine design), the blade will create an
impulsive noise every time it passes through the ‘wind shadow’ of the tower.
The various aerodynamic noise mechanisms are shown in Table 10.2 (Wagner et al.,1996). They
are divided into three groups: (1) low-frequency noise, (2) inflow turbulence noise, and (3) airfoil self
noise. A detailed discussion of the aerodynamic noise generation characteristics of a wind turbine is
beyond the scope of this work.

Table 13. 7 Wind turbine aerodynamic noise mechanisms (Wagner et al., 1996).

Type or indication Mechanism Main characteristics and importance


Low-frequency noise
Frequency is related to blade passing
Steady thickness noise; Rotation of blades or
frequency, not important at current
steady loading noise rotation of lifting surfaces
rotational speeds

Figure 11.6.3 Schematic of flow around a rotor blade. (Wagner et al.,


1996); U, wind speed.

| Environmental Aspects and Impacts Pavesi - 228 -


Frequency is related to blade passing
Passage of blades through
frequency, small in cases of upwind
Unsteady loading noise tower velocity deficit or
turbines/possibly contributing in case of
wakes
wind farms
Inflow turbulence Interaction of blades with Contributing to broadband noise; not yet
noise atmospheric turbulence fully quantified
Airfoil self-noise
Interaction of boundary
Broadband; main source of High-frequency
Trailing-edge noise layer turbulence with blade
noise (770 Hz <f <2 kHz)
trailing edge
Interaction of tip
Tip noise turbulence with blade tip Broadband; not fully understood
surface
Interaction of turbulence
Stall, separation noise Broadband
with blade surface
Non-linear boundary layer
Laminar boundary instabilities interacting
Tonal, can be avoided
layer noise with the
blade surface
Blunt trailing edge Vortex shedding at blunt
Tonal. can be avoided
noise trailing edge
Noise from flow over slits, vortex shedding from
holes. slits and Unstable shear flows over Tonal, can be avoided
intrusions holes and -intrusions

11.6.1.6 Mechanical noise

Mechanical noise originates from the relative motion of mechanical components and the dynamic
response among them. The main sources of such noise include:
 Gearbox
 Generator
 Yaw drives
 Auxiliary equipment (e.g., hydraulics)
 Cooling fans
Since the emitted noise is associated with the rotation of mechanical and electrical equipment, it
tends to be tonal (of a common frequency) in character, although it may have a broadband component.
For example, pure tones can be emitted from the rotational frequencies of shafts and generators, and the
meshing frequencies of the gears.

| Environmental Aspects and Impacts Pavesi - 229 -


In addition, the hub, rotor,
and tower may act as
loudspeakers, transmitting the
mechanical noise and radiating it.
The transmission path of the noise
can be air-borne (ah) or structure-
borne (sh). Air-borne means that
the noise is directly propagated
from the component surface or
interior into the air. Structure-
borne noise is transmitted along
other structural components before
it is radiated into the air. For
example, Figure 11.6.4shows the
type of transmission path and the
sound power levels for the
individual components determined
at a downwind position (1 15 m)
for a 2 MW wind turbine (Wagner
et al., 1996).
Note that the main source of
Figure 11.6.4 Components and total sound power level for wind mechanical noise was the gearbox
turbine (Wagner et al., 1996); LWA predicted sound power level; that radiated noise from the
ah, airborne; slb, structure borne. nacelle surfaces and the machinery
enclosure.

11.6.2 Noise prediction from wind turbines


11.6.2.1 Single wind turbines

The prediction of noise from a single wind turbine under expected operating conditions is an
important part of an environmental noise assessment. Considering the complexity of the problem, this
is not a simple task, and depending on the resources and time available can be quite involved. To
complicate matters, wind turbine technology and design has steadily improved through the years, so
prediction techniques based on experimental field data from operating turbines may not reflect state-of-
the-art machines.
Despite these problems, researchers have developed analytical models and computational codes
for the noise prediction of single wind turbines. In general, these models can be divided into the
following three classes (Wagner et al., 1996):
 Class 1. This class of models gives a simple estimate of the overall sound power level as a
function of basic wind turbine parameters (e.g., rotor diameter, power and wind speed).
They represent rules of thumb and are simple and easy to use
 Class 2. These consider the three types of noise mechanisms previously described, and
represent current turbine state-of-the-art
 Class 3. These models use refined models describing the noise generation mechanisms

| Environmental Aspects and Impacts Pavesi - 230 -


and relate them to a detailed description of the rotor geometry and aerodynamics
For the Class 1 models, there are empirical equations that have been used to estimate the sound
power level. For example, based on turbine technology up to about 1990, Bass (1993) states that a
useful rule of thumb is that wind turbines radiate 107 of their rated power as sound power. Examples of
three other Class 1 models for the prediction of sound power level are summarized in Equations
(10.9.4) to (10.9.6).
L WA  10  log10 PWR   50 (10.9.4)

LWA  22  log10 D   72 (10.9.5)

LWA  50  log10 VTip   10  log10 D   4 (10.9.6)

where LWA is the overall A-weighted sound power level, VTip is the tip speed at the rotor blade ( d
s ) , D is the rotor diameter (m), and PWT the rated power of the wind turbine (W).
The first two equations represent the simplest (and least accurate today, since they were
developed for older machines) methods to predict the noise level of a given turbine based on either its
rated power or rotor diameter. The last equation illustrates a rule of thumb that aerodynamic noise is
dependent on the fifth power or the tip speed.
The complexity of the Class 2 and 3 models is illustrated in Table 13. 8, where the input details
required for both models are summarized. A detailed discussion of all these models is beyond the scope
of this section and is given in the text of Wagner et al. ( 1996).

11.6.2.2 Multiple wind turbines

Intuitively, one would expect that doubling the number of wind turbines at a given location would
double the sound energy output, Since the decibel scale is logarithmic, the relation to use for the
addition of two sound pressure levels (L1 and L2) is given as:


Ltotal  10log10 10L1 10  10L2 10  (10.9.7)

This equation has two important implications:


 Adding sound pressure levels of equal value increases the noise level by 3 dB
 If the absolute value of L, - L, is greater than 15 dB, the addition of the lower level has
negligible effects
This relation can be generalized for N noise sources:
N
Ltotal  10log10 10Li 10 (10.9.8)
i 1

| Environmental Aspects and Impacts Pavesi - 231 -


Table 13. 8 Typical inputs for Class 2 and 3 noise prediction models Cj Vagner et al., 1996).

Class
Group Parameter Class 2
3
Turbine configuration Hub height X X
Type of tower (upwind or downwind) X

Blades and rotor Number of blades X X


Chord distribution (X) X
Thickness of trailing edge (X) X
Radius X X
Profile shape (X) X
Shape of blade tip (X) X
Twist distributor (X) X

Atmosphere Turbulence intensity X X


Ground surface roughness X X
Turbulence intensity spectrum X
Atmospheric stability conditions X

Turbine operation Rotational speed X X


Wind speed, alternatively: rated power, rated X X
wind speed, cut-in wind speed
Wind direction X

11.6.3 Noise propagation from wind turbines


In order to predict the sound pressure level at a distance from a source with a known power level,
one must consider how the sound waves propagate. Details of sound propagation in general are
discussed in Beranek and Ver (1992). For the case of a stand-alone wind turbine, one might calculate
the sound pressure level by assuming spherical spreading, which means that the sound pressure level is
reduced by 6 dB per doubling of distance. If the source is on a perfectly flat and reflecting surface,
however, then hemispherical spreading has to be assumed, which leads to a 3 dB reduction per
doubling of distance.
Furthermore, the effects of atmospheric absorption and the ground effect, both dependent on
frequency and the distance between the source and observer, have to be considered. The ground effect
is a function of the reflection coefficient of the ground and the height of the emission point.
Wind turbine noise also exhibits some special features (Wagner et al., 1996). First, the height of
the source is generally higher than conventional noise sources by an order of magnitude, which leads to
less importance of noise screening. In addition, the wind speed has a strong influence on the generated
noise. The prevailing wind directions can also cause considerable differences in sound pressure levels
between upwind and downwind positions.

| Environmental Aspects and Impacts Pavesi - 232 -


The development of an accurate noise propagation model generally must include the following
factors:
 Source characteristics (e.g., directivity, height, etc.)
 Distance of the source to the observer
 Air absorption
 Ground effect (i.e., reflection of sound on the ground, dependent on terrain cover, ground
properties, etc.)
 Propagation in complex terrain
 Weather effects (i.e., change of wind speed or temperature with height)
A discussion of complex propagation models that include all these factors is beyond the scope of
this work. A discussion of work in this area is given by Wagner et al. (1996). For estimation purposes,
a simple model based on hemispherical noise propagation over a reflective surface, including air
absorption, is given as:
Lp LW  10log10  2R 2   R (10.9.9)

where Lp is the sound pressure level (dB) a distance R from a noise source radiating at a power
level LW (dB) and  is the frequency-dependent sound absorption coefficient.
This equation can be used with either broadband sound power levels and a broadband estimate of
the sound absorption coefficient [ = 0.005 dB(A) m-1 or more preferably in octave bands using octave
band power and sound absorption data.

11.6.4 Noise reduction methods for wind turbines


Turbines can be designed or retrofitted to minimize mechanical noise. This can include special
finishing of gear teeth. using low-speed cooling fans, mounting components in the nacelle instead of at
ground level, adding baffles and acoustic insulation to the nacelle, using vibration isolators and soft
mounts for major components, and designing the turbine to prevent noises from being transmitted into
the overall structure. Furthermore, if low frequency noise is a problem in an area, the permitting agency
may prohibit the installation of downwind machines (NWCC, 1998).
If a wind turbine has been designed using appropriate design procedures (as described in Chapter
4). it is likely that new noise reducing airfoils will have been used, and mechanical noise emissions will
not be a problem. In general, designers trying to reduce wind turbine noise even more must concentrate
on the further reduction of aerodynamic noise. It has been previously noted that the following three
mechanisms of aerodynamic noise generation are important for wind turbines (assuming that tonal
contributions due to slits, holes, trailing edge bluntness, control surfaces, etc., can be avoided by proper
blade design):
 Trailing edge noise
 Tip noise
 Inflow turbulence noise
A review of work in these three areas is beyond the scope of this text and here the reader is
referred to the text of Wagner et al. (1996). It should be noted that noise has been reduced in modern
turbine designs via the use of lower tip speed ratios, lower blade angle of attack, upwind designs, and,
most recently, by using specially modified blade trailing edges.

11.6.5 Noise standards or regulations

| Environmental Aspects and Impacts Pavesi - 233 -


An appropriate noise assessment study should contain the following three major parts of
information:
 A survey of the existing ambient background noise levels
 Prediction (or measurement) of noise levels from the turbine(s) at and near the site
 An assessment of the acceptability of the turbine(s) noise level
At the present time,
Table 13. 9 Noise limits of equivalent sound pressure levels, L, there are no common
[dB(A)]: European countries (Gipe, 1995) international noise
Country Commercial Mixed Residential Rural standards or regulations,
Denmark 40 45 especially ones that pertain
Germany to all of the above
day 65 60 55 50 information. In most
night 50 45 40 35 countries, however, noise
Netherlands regulations define upper
day 50 45 40 bounds for the noise to
night 40 35 30 which people may be
exposed. These limits
depend on the country and are different for daytime and night time. In Europe, as shown in Table 13. 9,
fixed noise limits are the standard (Gipe, 1995).
In the United States, although no formal federal noise regulations exist, the US Environmental
Protection Agency (EPA) has established noise guidelines. Many states do have noise regulations, and
many local governments have enacted noise ordinances.
Examples of such ordinances for wind turbines are given in the latest edition of Permitting of
Wind Energy Facilities: A Handbook (NWCC, 1998).
It should also be pointed out that imposing a fixed noise level standard might not prevent noise
complaints. This is due to the question of relative level of broadband background turbine noise
compared to changes in background noise levels (NWCC, 1998). That is, if tonal noises are present,
higher levels of broadband background noise axe needed to effectively mask the tone(s). Accordingly,
it is common for community noise standards to incorporate a penalty for pure tones, typically 5 dB(A).
Therefore, if a wind turbine meets a sound power level standard of 45 dB(A), but produces a strong
whistling, 5 dB(A) are subtracted from the standard. This forces the wind turbine to meet a real
standard of 40 dB(A).
A discussion of noise measurement techniques that are specific to wind turbine standards or
regulations is beyond the scope of this text. A review of such techniques is given in Hubbard and
Shepherd (1990), Germanisher Lloyd (1994), and Wagner et al. (1996).

11.7 Electromagnetic Interference Effects


Wind turbines can present an obstacle for incident electromagnetic waves, which may be
reflected, scattered, or diffracted by the wind turbine. As shown schematically in Figure 11.7.1
(Wagner et al., 1996), when a wind turbine is placed between a radio, television, or microwave
transmitter and receiver, it can sometimes reflect portions of the electromagnetic radiation in such a
way that the reflected wave interferes with the original signal arriving at the receiver. This can cause
the received signal to be significantly distorted.
Some key parameters that influence the extent of electromagnetic interference (EMI) caused by
wind turbines include:

| Environmental Aspects and Impacts Pavesi - 234 -


Figure 11.7.1 Scattering of electromagnetic signals by a wind
turbine (Wagner et al., 1996).
 Type of wind turbine (i.e., horizontal axis wind turbine (HAW) or vertical axis wind
turbine (VAWT))
 Wind turbine dimensions
 Turbine rotational speed
 Blade construction material
 Blade angle and geometry
 Tower geometry
In practice, the blade construction material and rotational speed are key parameters. For example,
older HAWS with rotating metal blades have caused television interference in areas near the turbine.
Today EM1 from wind turbines is less likely because most blades are now made from composite
materials. Most modern machines, however, have lightning protection on the blade surfaces, which can
increase electromagnetic interference.

11.7.1 Characterization of electromagnetic interference from wind turbines


11.7.1.1 Mechanism of electromagnetic interference

Electromagnetic interference from wind turbines is generated from multiple path effects.
That is, as shown in Figure 11.7.2, in the vicinity of a wind installation, two transmission paths
occur linking the radio transmitter to the receiver. Multiple paths occur on many radio transmissions
(caused by large buildings or other structures). The feature unique to wind turbines, however, is blade

| Environmental Aspects and Impacts Pavesi - 235 -


rotation which causes changes, over short time intervals, in the length of the secondary or scattered
path. Thus, the receiver may acquire two signals simultaneously, with the secondary signal causing
EM1 because delay or distortion of the signal varies with time.
Effects of the signal variation change depend on the modulation scheme, that is, the manner in
which the received information is coded. For example, for amplitude modulated (AM) signals, the
variation in signal level can be highly undesirable. Interference may also be produced in frequency
modulated (FM) signal. Furthermore, the doppler shift caused by rotation of the blades may interfere
with radar, and for digital systems, the signal variation can increase the bit error rate (Chignell, 1986).

11.7.1.2 Wind turbine parameters

A wind turbine may provide a number of different electromagnetic scattering mechanisms.


The turbine blades in particular can play a significant role in generating electromagnetic
interference. They may scatter a signal directly as they rotate and they may also scatter signals reflected
from the tower. The degree of EM1 caused by wind turbines is influenced by numerous factors,
including:
 The wide frequency range of radio signals
 The variety of modulation schemes
 The wide variation in wind turbine parameters
There are a multitude of possible ways whereby a wind turbine can modulate the radio signal and
cause interference. Situations have either been reported or postulated where almost every design
parameter of the wind turbine system may be critical to a specific radio service. The following general
comments on the most important design variables can be made (Chignell, 1986):
Type of machine Different waveforms have been observed in the interference generated by
horizontal and vertical axis machines. Most work in this area has concentrated on television
interference (TVI) effects.
Machine dimensions The overall dimensions, particularly the diameter of the rotor, are
important for establishing the radio frequency bands where interference may occur. Specifically, the
larger the machine, the lower the frequency above which radio services may be affected. That is, a
large machine will affect HF, VHF, UHF, and microwave bands, while a small machine may degrade
only UHF and microwave transmissions.
Rotational speed The rotational speed of the wind turbine and the number of blades determine
the modulation frequencies in the interfering radio or telecommunications signal. If one of these
coincides with a critical parameter in the radio or telecommunications receiver, the interference is
increased.
Blade Construction The blade cross-section and material can be significant here. For example,
the following general observations have been made:
 The geometry of the blade should be simple; ideally a combination of simple curves
which avoid sharp corners and edges
 In general, scattering from fibreglass or wooden blades is less than from a comparable
metal structure. Fibreglass is partially transparent to radio waves whereas wood absorbs
them, not allowing energy to be scattered
 The addition of any metallic structure such as a lightning conductor or a metal blade root
to a fibreglass blade may negate the material advantages, with the composite structure
scattering more effectively than an all-metal blade. To avoid this problem, the metallic

| Environmental Aspects and Impacts Pavesi - 236 -


components should avoid sharp edges and corners
Geometry In larger machines, and with microwave signals, the angles defining the area where
interference occurs become so narrow that small changes in the yaw, twist, and pitch of the blades
become important along with the tilt, cone, and teeter angles.
Tower The tower may also scatter radio waves and as the wind machine rotates, the blades can
‘chop’ this signal. The geometry of the tower should be kept simple, but if a complicated lattice is
essential, a careful choice of angles may reduce the problem. Note, however, that Sengupta and Senior
(1994) have not found the tower to be a significant source of EMI.

11.7.2 Potential electromagnetic interference effects of wind turbines


For electromagnetic interference generated by a wind turbine to disturb a radio signal, the
following conditions must be satisfied:
 A radio transmission must be present
 The wind system must modify the radio signal
 A radio receiver must be present in the volume affected by the wind system
 The radio receiver must be susceptible to the modified signal
Based on EM1 experiments, field experience, and analytical modelling in the United States
(Sengupta and Senior, 1994) and Europe (Chignell, 1986), a summary of the effects of wind turbine
EM1 on various radio transmissions is as follows:
 Television interference Most reports of electromagnetic interference from wind turbines
concern television service. Television interference from wind turbines is characterized by
video distortion that generally occurs in the form of a jittering of the picture that is
synchronized with the blade passage frequency (rotor speed times number of blades). A
significant amount of work on this subject has been carried out in the United States and
Europe to quantify this effect
 FM radio interference Effects on FM broadcast reception have only been observed in
laboratory simulations. They appear in the form of a background ‘hiss’ superimposed on
the FM sound. This work concluded that the effects of wind turbine EM1 to FM reception
was negligible except possibly within a few tens of meters from the wind turbine
 Interference to aircraft navigation and landing systems The effects on VOR (VHF
omnidirectional ranging) and LORAN (a long range version of VOR) systems have been
studied via analytical models. Results from the VOR studies indicate that a stopped wind
turbine may produce errors in the navigational information produced by the VOR. When
the wind turbine is operating, however, the potential interference effects are significantly
reduced. Note that existing Federal Aviation Authority (FAA) rules prevent a structure the
size of many wind turbines being erected within 1 km of a VOR station. For LORAN
systems, which operate at very low frequencies, no degradation in communication
performance is likely to occur. This assumes that a wind turbine is not in close proximity
to the transmitter or receiver
 Interference to microwave links Analytical work has indicated that electromagnetic
interference effects tend to smear out the modulation used in typical microwave
transmission systems. In Europe, experimental work has produced reports of major
interference problems on microwave links
 Interference with cellular telephones. Since cellular radio is designed to operate in a

| Environmental Aspects and Impacts Pavesi - 237 -


mobile environment, it should be comparatively insensitive to EM1 effects from wind
turbines
 Interference with satellite services. Satellite services using a geostationary orbit are not
likely to be affected because of the elevation angle in most latitudes and the antenna gain

11.7.3 Prediction of electromagnetic interference effects from wind turbines: analytical models
A summary of the most detailed and general models for the analysis of radio signals with
electromagnetic interference from wind turbines is given by Sengupta and Senior (1994). In this report
the authors summarize over 20 years of work directed toward this subject focusing on large-scale wind
turbine systems. Specifically, they have developed a general analytical model for the mechanism by
which a wind turbine can produce electromagnetic interference. Their model is based on the schematic
system shown in Figure 11.7.2, which illustrates the field conditions under which a wind turbine can
cause EMI. As shown, a transmitter, T, sends a direct signal to two receivers, R, and a wind turbine,
WT. The rotating blades of the wind turbine both scatter and transmit a scattered signal. Therefore, the
receivers may acquire two signals simultaneously, with the scattered signal causing the EM1 because it
is delayed in time or distorted. Signals reflected in a manner analogous to mirror reflection are defined
as back-scattered (about 80% of the region around the wind turbine). Signal scattering that is analogous
to shadowing is called forward-scattering and represents about 20% of the region around a turbine.
For more details of this analysis the reader is referred to Sengupta and Senior (1994), where
analytical expressions are developed for the signal power interference (expressed in terms of a
modulation index), and the signal scatter ratio, important EMI parameters. This work also summarizes
their analytical scattering models for various wind turbine rotors.
Their approach here was to develop simplified, idealized models of HAWT and VAWT rotors
and to compare the model predictions of signal scattering with measured scattering.
This approach is used to illustrate the basic principles of EM1 from wind turbines and to provide
useful equations for estimating the magnitude of potential interference in practical situations. An
example of this is contained in their reports designed to assess TV interference from large and small
wind turbines (Senior and Sengupta, 1983; Sengupta et al., 1983).

11.7.4 Simplified analysis


At the present time, the general, and even specialized, models developed by US researchers that
have been previously described are not readily usable for wind system designers.
Simplified models, however, can be used to predict potential EM1 problems from a wind turbine.
For example, Van Kats and Van Rees (1989) developed a simple EM1 interference model for wind
turbines that could be used to predict their impact on TV broadcast reception. As summarized below,
results from this model produce a calculated value for the signal-to-interference ratio (C/Z) for UHF
TV broadcast reception in the area around the wind system that can used to evaluate TV picture quality.
The general geometry and definition of terms for the model is given below in Figure 11.7.2. This
model is based on the assumption that the wind turbine behaves as an obstacle for incident
electromagnetic waves, serving as a secondary source for radiation.
In this model, the spreading of this secondary field over its environment is determined by the
following factors:
 Size and shape of the obstacle with respect to signal wavelength, 
 Dielectric and conductive properties of components of the wind turbine

| Environmental Aspects and Impacts Pavesi - 238 -


 Position of the blades and the structure of the wind turbine with respect to the polarization
of the incident wave
Because the wind turbine blades are in continuous motion, the impact of the reflected and
scattered fields is complex. Thus, the calculation of these fields requires empirical correlation. To
approximate the behaviour of the system the model is based on radar technology, where an obstacle is
characterized by its (bistatic) radar cross-section, b.
(This variable is defined as the imaginary surface of an isotropic radiator from which the received
power corresponds to the actual power received from the obstacle.) In general, ob is a function of the
dielectric properties and geometry of the obstacle and the signal wavelength.
When the impact of a wind
turbine on radio waves is
modelled using b, an expression
for the signal-to-interference ratio,
C/l, can be found in the area
around the wind turbine. It should
be noted that for each specific
radio service (e.g., TV, mobile
radio, microwave links, etc.) a
different C/I ratio will be required
for reliable operation. Using the
assumed geometry and
nomenclature shown in Figure
11.7.2, the C/l ratio at a distance r
from the turbine can be calculated
by separately calculating the
strength (in dBW) of the desired
signal, C, and the strength of the Figure 11.7.2 Model configuration for electromagnetic
interference signal, I. The desired interference due to a wind turbine (Sengupta and Senior, 1994).
signal is given by:
C = Pt + (L1 + A1) + Gr (10.9.10)
where Pl is the transmitter power, L1 is free space path loss between transmitter and the receiver, A1 is
the additional path loss between transmitter and the receiver, and Gr receiver gain. In a manner similar
to acoustic measurements, signal strength (in dBW) is expressed in decibels (here, dBW or dB above 1
W). The undesired signal, I, is given by:
 4 
I  Pl   L2  A2   10log  2 b    Lr  Ar   G 'r (10.9.11)
  
where L2 is free space path loss between the transmitter and the obstacle, Ar is the additional path loss
between the transmitter and the obstacle, Lr is free space path loss between the obstacle and the
receiver, Ar is the additional path loss between the obstacle and the receiver, and Gr’ is the receiver
gain (obstacle path). Next, assume that L1 = L when the distance of the wind turbine and the receiver to
the TV transmitter is large. In this equation, the free space loss, Lr (in dB), can be calculated from:
Lr  20log  4   20log  r   20log    (10.9.12)

| Environmental Aspects and Impacts Pavesi - 239 -


If one defines the antenna discrimination factor, G, as:
AG = Gr –Gr’ (10.9.13)
Then, the C/l ratio (in dB) becomes:
C
 10log  4   20log  r   10log  b   A 2  A1  A r  G (10.9.14)
I
From this equation, it follows that C/I can be improved by:
 Decreasing b (reducing the effect of the wind turbine)
 Increasing G (better directivity of the antenna)
 Increasing the additional loss A2 + Ar on the interference path
 Decreasing the additional loss A1 in the direct path
The control of these parameters depends strongly on the location of the receiver to the wind
system. In general, two alternative situations can be distinguished
 A significant delay time, t, between the desired signal (C) and the undesired signal (I).
The receiver is located in an area between the wind turbine and the transmitter. In this area
the undesired signal is dominated by reflection and scatter from the wind turbine (back-
scatter region).
 No delay time between the desired signal (C) and the undesired signal (I). This
phenomenon occurs in places where the receiver is behind the wind turbine, as seen from
the transmitter. In this area the undesired signal can only be the result of scatter or
refraction at the wind turbine (forward-scatter region).
In Equation (10.9.14), the distance r from the wind turbine to the receiver can be considered as a
variable. When the factors A1, A2, and Ar are assumed constant, G = 0, and b of the wind turbine is
known, then r = f(C/I) can be calculated.
When C/I is defined as that signal-to-interference ratio required for a specific radio service, a
curve with radius r around a wind turbine can be drawn. This will define an area within which the
required C/I will not be satisfied. Thus, using this method, criteria for the sitting of a particular design
of wind turbine with respect to its electromagnetic interference on TV reception can be produced,
provided that:
 The C/I required for TV reception is known (experimentally determined examples of this
ratio are presented in the original reference)
 The bistatic cross-section, b , of the wind turbine is known (again, experimentally
determined examples of this parameter for both the forward- and back-scattering regions
are presented in the original reference)

11.7.5 Resources for estimation

| Environmental Aspects and Impacts Pavesi - 240 -


As can be seen from the previous discussion and as noted by Chignell (1986), at present, it is
impossible to provide a wind system designer with complete technical guidance on the subject of EMI.
Thus, problems have to be addressed on a site-by-site basis. In Europe, the International Energy
Agency (IEA) Expert Group has recommended an interim assessment procedure for identifying when
electromagnetic interference may arise at a particular site.
This procedure can be regarded as a means of warning when a problem may arise, but, at present,
it does not make a recommendation for dealing with such situations. For an initial step, one must
establish what radio services are present in the volume occupied by the wind system - this means
identifying the radio transmitters. Ideally, in each location there would be a central register of all
transmitters that the wind system designer could access, but this is not usually the case. The IEA
recommends that this should include approaches to radio regulatory agencies, visual observations at the
site and on appropriate maps, and a site survey to monitor the radio transmissions that are actually
present. Also, the survey should be long enough to include services that are only present on a part-time
basis and should note the requirements of mobile users including emergency services, aircraft,
shipping, and public utilities.
Once the radio services present have been identified and the transmitters have been marked on a
map, the interference zones should be determined. This may be the most complex part of the process,
as only the detailed analytical or experimental measurement techniques described by Sengupta and
Senior (1994) are presently available. The next part of the procedure is to determine if any receivers for
that service will be present in the interference zone. For television broadcasting, this essentially means
establishing if there are any dwellings in the zones. If receivers exist, further advice should be sought.

Figure 11.7.3 Geometry for interference of the electromagnetic path by an


obstacle (Van Kats and Van Rees, 1989).

| Environmental Aspects and Impacts Pavesi - 241 -


Note that the presence of such a receiver does not necessarily mean that EM1 problem will arise – it
also depends on whether the radio service is robust to major changes in its signal level.

11.8 Land-use environmental impacts


There are a number of land-use issues to be considered when sitting wind turbines, Some of them
involve government regulations and permitting (such as zoning, building permits, and approval of
aviation authorities). Others may not be subject to regulation, but do have an impact on public
acceptance. The following are some of the major land-use issues:
 Actual land required per energy output or capacity per unit of land area
 Amount of land potentially disturbed by a wind farm
 Non-exclusive land-use and compatibility
 Rural preservation
 Turbine density
 Access roads and erosion and/or dust emissions
It is beyond the scope of this text to go into the extensive details of these factors. Gipe (1995)
present an overview of them (and other land-use considerations). The next sections presents a summary
of the most general land-use considerations for wind energy systems, and potential strategies that can
be used to minimize the environmental problems associated with wind turbine land-use.

11.8.1 Land-use considerations


Compared to other power plants, wind generation systems are sometimes considered to be land
intrusive rather them land intensive. The major intrusive effects, via visual impacts, were addressed in
Section 11.3. On the extent of land required per unit of power capacity, wind farms require more land
than most energy technologies. However, while wind energy system facilities may extend over a large
geographic area, the physical ‘footprint’ of the actual wind turbine and supporting equipment only
covers a small portion of the land. In the United States, for example, wind farm system facilities may
occupy only three to five percent of the total acreage, leaving the rest available for other uses. In
Europe, it has been found that the percentage of land-use by actual facilities is less than the US wind
farms in California. For example, U.K. wind farm developers have found that typically only 1 % of the
land covered by a wind farm is occupied by the turbines, substations, and access roads.
Also, in numerous European projects, farm land is cultivated up to the base of the tower.
When access is needed for heavy equipment, temporary roads are placed over tilled soil. Thus,
European wind farms only occupy from one to three percent of the available land.
When one determines the actual amount of land used by wind farm systems, it is also important to
note the influence of the wind turbine spacing and placement. Wind farms can occupy Erom 4 to 32
hectares (10 to 80 acres) per megawatt of installed capacity. The dense arrays of the California wind
farms have occupied from about 6 to 7 hectares (15 to 18 acres) per megawatt of installed capacity.
Typical European wind farms have the wind turbines spread out more and generally occupy 13 to 20
hectares (30 to 50 acres) per megawatt of installed capacity (Gipe, 1995).
Since wind generation is limited to areas where weather patterns provide consistent wind
resources over a long season, the development of wind power in the US has occurred primarily in r u d
and relatively open areas. These lands are often used for agriculture, grazing, recreation, open space,
scenic areas, wildlife habitat, and forest management. Wind development is generally compatible with

| Environmental Aspects and Impacts Pavesi - 242 -


the agricultural or grazing use of a site. Although activities in these areas may be disrupted during
construction, only intensive agricultural uses may be reduced or modified during the project’s operation
(NWCC, 1998). In Europe, owing to higher population densities, there are many competing demands
for land, and wind farms have tended to be of a smaller total size.
The development of a wind farm may affect other uses on or adjacent to a site. For example,
some parks and recreational uses that emphasize wilderness values and reserves dedicated to the
protection of wildlife (e.g., birds) may not be compatible with wind farm development. Other uses,
such as open space preservation, growth management, or non wilderness recreational facilities may be
compatible depending on set-back requirements, the nature of on-site development, and the effect on
resources of regional importance (NWCC, 1998).
In general, the variables that may determine land-use impacts include:
 Site topography
 Size, number, output, and spacing of wind turbines
 Location and design of roads
 Location of supporting facilities (consolidated or dispersed)
 Location of electric lines (overhead, or underground)

11.8.2 Mitigation of land-use problems


A wide range of actions is available to ensure that wind energy projects are consistent and
compatible with most existing and planned land uses. Many of these involve the layout and design of
the wind farm. For example, where wind energy development is located in or near recreational or
scenic open space uses, some permitting agencies have established requirements intended to ‘soften’
the industrial nature of the project. As summarized by the National Wind Coordinating Committee
(NWCC, 1998), these include the following (again, note that many are a result of visual impact
considerations):
 Selecting equipment with minimal structural supports, such as guy wires
 Requiring electrical collection lines to be placed underground
 Requiring maintenance facilities to be off-site
 Consolidating equipment on the turbine tower or foundation pad
 Consolidating structures within a wind farm area
 Requiring the use of more efficient or larger turbines to minimize the number of turbines
required to achieve a specific level of electrical output
 Selecting turbine spacing and types to reduce the density of machines and avoid the
appearance of ‘wind walls’
 Use of roadless construction and maintenance techniques to reduce temporary and
permanent land loss
 Restricting most vehicle travel to existing access roads
 Limiting the number of new access roads, width of new roads, and avoiding or
minimizing cut and fill
 Limiting placement of turbines and transmission towers in areas with steep, open
topography to minimize cut and fill
From a land-use viewpoint, this complete list is really a goal for a ‘perfect’ wind site, and
permitting agencies should consider the following points when determining which of these should be
applied to a particular site:

| Environmental Aspects and Impacts Pavesi - 243 -


 Cost associated with a particular strategy
 Type and level of impact
 Land-use objectives of the community
 Significance of any potential land-use inconsistency or incompatibility
 Available alternatives
Many of the previous requirements have been used in wind energy projects in Europe. Here, such
projects are often located in rural or agricultural areas, and have tended to consist of single dispersed
units or small clusters of wind turbines. In several European countries, wind turbines have been placed
on dikes or levees along coastal beaches and jetties, or just off shore. Wind turbines sited in farming
areas have also been located to minimize disturbances to planting patterns, and the permanent
subsystems have been consolidated within a single right of way on field edges, in hedgerows, or along
farm roads. Often, no access roads have been constructed and erection or periodic maintenance is
carried out with moveable cushioned mats or grating.
In the United States, other land-use strategies associated with wind farm sites include the use of
buffer zones and setbacks to separate the project from other potentially sensitive or incompatible land
uses. The extent of this separation varies, depending on the area’s land use objectives and other
concerns such as visual aspects, noise, and public safety. Also, on some wind farm projects, the
resource managers for adjacent properties have established lease or permit conditions such that wind
farm development on one site does not block the use of the wind resource on adjacent sites.

11.9 Other Environmental Considerations


This section will summarize some areas that should be considered when assessing the
environmental impact of a wind energy system. These include safety, general impacts on flora and
fauna, and shadow flicker.

11.9.1 Safety
Safety considerations include both public safety and occupational safety. Here, the discussion will
be centered on the public safety aspects (although some occupational safety issues will be included as
well). A review of occupational safety in the wind industry is contained in the work of Gipe (1995).
In the public safety area, the primary considerations associated with wind energy systems are
related to the movement of the rotor and the presence of industrial equipment in areas that are
potentially accessible to the public. Also, depending on the site location, wind energy system facilities
may also represent an increased fire hazard. The following aspects of public safety are important
(NWCC, 1998):
 Blade throw. One of the major safety risks from a wind turbine is that a blade or blade
fragment can be thrown from a rotating machine. Wind turbines that have guy wires or
other supports can also be damaged. Turbine nacelle covers and rotor nose cones can also
blow off machines. In actuality, these events are rare and usually occur in extreme wind
conditions, when other structures are susceptible. The distance a blade, or turbine part,
may be thrown depends on many variables (e.g., turbine size, height, size of broken part,
wind conditions, topography, etc.) and rarely has exceeded 500 m, with most pieces found
within 100 to 200 m of the tower. A detailed example of a blade throw calculation is given
by Turner (1989)
 Falling ice or thrown ice. Safety problems can occur when low temperatures and

| Environmental Aspects and Impacts Pavesi - 244 -


precipitation cause a build-up of ice on turbine blades. As the blades warm, the ice melts
and either falls to the ground, or can be thrown from the rotating blade. Falling ice from
nacelles or towers can also be dangerous to people directly under the wind turbine. A
detailed technical review of the safety aspects of this problem is summarized by Bossanyi
and Morgan (1996)
 Tower failure. The complete failure of wind turbine towers or guy wires usually brings
the entire wind turbine to the ground if the rotor is turning, or if the problem is not
detected immediately. High ice loads, poor tower or foundation design, corrosion, and
high winds can increase this potential safety risk
 Attractive nuisance. Despite their usual rural location, many wind energy system sites
are visible from public highways, and are relatively accessible to the public. Because the
technology and the equipment associated with a wind turbine site are new and unusual, it
can be an attraction to curious individuals. Members of the public who attempt to climb
towers or open access doors or electrical panels could be subject to injuries from moving
equipment during operation, electrical equipment during operation, or numerous other
hazardous situations
 Fire hazard. In arid locations, site conditions that are preferable for development of wind
sites such as high average wind speeds, low vegetation, few trees, etc., may also pose a
high fire hazard potential during the dry months of the year. Particularly vulnerable sites
are those located in rural areas where dry land grain farming occurs or the natural
vegetation grows uncontrolled and is available for fuel. In these types of locations fires
have started from sparks or flames as a result of numerous causes, including: substandard
machine maintenance, poor welding practices, electrical shorts, equipment striking power
lines, and lightning
 Worker hazard. For any industrial activity, there is the potential for injury or the loss of
life to individuals. At the present time there are no statistics that can be used to compare
wind facility work with work at othes energy-producing facilities. There have, however,
been several fatalities associated with wind turbine installation and maintenance (see
Gipe, 1995)
 Electromagnetic fields. Electrical and magnetic fields are caused by the flow of electric
current through a conductor, such as a transmission line. The magnetic field is created in
the space around the conductor, and its field intensity decreases rapidly with distance. In
recent years, some members of the public have been concerned regarding the potential for
health effects associated with electromagnetic fields

11.9.2 Mitigation of public health safety risks


The following summarizes how the public health safety risks previously mentioned can be
mitigated:
 Blade throw. The most appropriate method for reducing the blade throw potential is the
application of sound engineering design combined with a high degree of quality control.
Today, it is expected that blade throw should be rare. For example, braking systems, pitch
controls, and other speed controls on wind turbines should prevent design limits from
being exceeded. Because of safety concerns with blade throw and structural failure, many
permitting agencies have separated wind turbines from residences, public travel routes,
and other land uses by a safety buffer zone or setback. Examples of typical regulations for

| Environmental Aspects and Impacts Pavesi - 245 -


the United States are contained in the permitting handbook of the National Wind
Coordinating Committee (NWCC, 1998)
 Falling ice or thrown ice. To reduce the potential for injury to workers, discussion of
blade throw and ice throw hazards should be included in worker training and safety
programs. Also, project operators should not allow work crews near the wind turbines
during very windy and icing conditions
 Tower failure. Complete failure of the structure should be unlikely for turbines designed
according to modern safety standards. Additional security may be obtained by locating the
turbine at least the height of the tower away from inhabited structures
 Attractive nuisance. Many jurisdictions have required fencing and posting at wind
project boundaries to prevent unauthorized access to the site. However, other jurisdictions
prefer that the land remains unfenced, particularly if located away from well-traveled
public roads, so that the area appears to remain open and retains a relatively natural
character. Many jurisdictions require the developer to post signs with a 24-hour toll-free
emergency phone number at specified intervals around the perimeter, if the area is fenced,
and throughout the wind site, if it is unfenced. Also, liability concerns dictate that access
to towers and electrical equipment be locked, and that warning signs be placed on towers,
electrical panels, and at the project entrance
 Fire hazard. The single most effective fire hazard avoidance measure is to place all
wiring underground between the wind turbines and the project substation. In fire-prone
areas, most agencies establish permit conditions for the project which address the potential
for fire hazard, and a fire control plan may be required. In addition, most agencies require
fire prevention plans and training programs to further reduce the potential for fires to
escape the project and spread into surrounding areas
 Worker hazard. To minimize this hazard, the wind energy project should follow well
established worker protection requirements for the construction and operation of wind
energy sites
 Electromagnetic fields. Given the rural nature of most wind farm sites, and the relatively
low power levels transmitted in a typical site, this safety problem may not represent a
major public safety hazard. In the United States, a few states have sought to limit field
exposure levels to the levels from existing transmission lines by specifying limits on field
strengths, either within or at the edge of the rights-of-way for new lines

11.9.3 Impact on flora and fauna


Since wind turbine sites are typically located in rural areas that are either undeveloped or used for
grazing or farming, they have the potential to directly and indirectly affect biological resources. For
example, since the construction of wind farms often involves the building of access roads and the use
of heavy equipment it is most probable that there will be some disturbance of the flora and fauna during
this phase of the project.
It should also be noted that the biological resources of concern include a broad variety of plants
and animals that live, use, or pass through an area. They also include the habitat that supports the living
resources, including physical features such as soil and water, and the biological components that
sustain living communities. These range from bacteria and fungi through the predators on top of the
food chain.
The effects on, or conflicts with, flora and fauna, if any, will depend on the plants and animals

| Environmental Aspects and Impacts Pavesi - 246 -


present, and the design and layout of the wind energy facility. In some cases, permitting agencies have
discouraged or prevented development because of likely adverse consequences on these resources. In
cases where sensitive resources are not present, or where adverse effects could be avoided or mitigated,
development has been permitted to proceed.

11.9.4 Mitigation measures


An important consideration regarding the flora and fauna that are found at a site is the potential
loss of habitat. Many species are protected state or national laws. It may therefore be useful to consult
with ecological or biological specialists early in the planning stages of a project, in order to make sure
that sensitive areas are not disturbed, and so that suitable mitigation measures are taken. Any survey
work should be carried out at the appropriate time of the year in order to take account of the seasonal
nature of some of the potential biological effects.
The wind system developer needs to meet with the local zoning or planning authority, and
relevant ecological/biological consultants, in order to discuss the timing of construction and the
placement of wind turbines and access roads to avoid at-risk species or habitats. In addition, there may
be requirements for on-going monitoring or an overall ecological management plan during the
construction period and during the lifetime of the wind system.

11.9.5 Shadow flicker and pashing


Shadow flicker occurs when the moving blades of the wind turbine rotor cast moving shadows
that cause a flickering effect, which could annoy people living close to the turbine.
In a similar manner, it is possible for sunlight to be reflected from gloss-surfaced turbine blades
and cause a ‘flashing’ effect. This effect has not surfaced as a real environmental problem in the United
States. It has been more of a problem in Northern Europe, because of the latitude and the low angle of
the sun in the winter sky, and because of the close proximity between inhabited buildings and wind
turbines.
For example, Figure 11.9.1 shows the results of calculations to determine the annual duration of
the shadow flicker effect for a location in Denmark (European Commission, 1998). In this figure there
are two houses marked as A and B that are respectively 6 and 7 hub heights from the turbine in the
centre of the diagram. This figure shows that house A will have a shadow from the turbine for 5 hours
per year. House B will have a shadow for about 12 hours per year. Note that the results €or this type of
calculation vary for different geographical regions due to different allowances for cloud cover and
latitude.

| Environmental Aspects and Impacts Pavesi - 247 -


Figure 11.9.1 Diagram of shadow flicker calculation (European
Commission, 1998); A, B howes

11.9.6 Mitigation measures


In the worst cases, flickering will only occur for a short duration (on the order of 30 minutes a
day for 10-14 weeks of the winter season). In Europe, one proposed solution is to not run the turbines
during these short time periods, while another is to site machines carefully, taking account of the
shadow path on nearby residences. Flashing can be prevented by the use of non-reflective, non-gloss
blades, and also by careful sitting.
A common guideline used in Denmark is to have a minimum distance of 6 to 8 rotor diameters
between the wind turbine and the closest neighbour. Houses located at a distance of 6 rotor diameters
(about 250 meters for a 600 kW machine) from a wind turbine in any of the sectors shown in Figure
10.11 will be affected for two periods each of 5 weeks duration per year (European Commission,
1998).

| Environmental Aspects and Impacts Pavesi - 248 -


11.10 References
Ackerman, T., Soder, L. (2000) Wind Energy Technology and Current Status: A Review, Renewable
and Sustainable Energy Reviews, 4, 3 15-374.
Allam RJ, Spilsbury CG (1992) A study of the extraction of CO2 from the flue gas of a 500 MW
pulverized coal fired boiler. Proceedings of the 1st International Conference on Carbon Removal
(FICCDR), Energy Conversion and Management, Pergoman Press
Anderson R, Morrison M, Sinclair K, Strickland D (1999) Studying wind energy/bird interactions: A
guidance document. Metrics and methods for determining or monitoring potential impacts of
birds at existing and proposed wind energy sites. National Wind coordination committee
(NWCC), Washington DC
Anderson, R. L.. Kendall, W., Mayer, L. S., Morrison, M. L., Sinclair, K.. Strickland, D., Ugoretz, S.
(1997) Standard Metrics and Methods for Conducting Avian Wind Energy Interaction Studies,
Proc. Windpower '97, AWEA, 265-272.
Aubrey, C. (2000) Turbine Colors.. .Do They Have to be Grey? Wind Directions, March, 18-19.
AWEA (1989) Procedure for measurement of acoustic emissions from wind turbine generator systems.
TeirI – standard 2.1, American Wind Energy Association
AWEA (2000) Facts about wind energy and noise. American Wind Energy Association, Washington
D.C.
AWEA (2002) The most frequently asked questions about wind energy. American Wind Energy
Association, Washington D.C
Bass, J. H. (1993) Environmental Aspects - Noise, in course notes for Principles of Wind Energy
Conversion, ed. L. L. Freris, Imperial College, London.
Benner JBH (1992) Impact of wind turbines on bird life: An overview of existing data and lacks in
knowledge inorder of the European community. Final report, Consultants on Energy and
Environment (CEE), Rotterdam, The Netherlands
Beranek, L. L., Ver, I. L. (1992) Noise and Vibration Control Engineering: Principles and Applications,
Wiley, New York.
Bossanyi, E. A., Morgan, C. A. (1996) Wind Turbine Icing - Its Implications for Public Safety, Proc.
I996 European Union Wind Energy Conference, 160-164.
Chignell, R. J. (1986) Electromagnetic Interference from Wind Energy Conversion Systems –
Preliiminary Information, Proc. European Wind Energy Conference 'SS, 583-586.
Clausager, I., Nohr, H. (1996) Impact of Wind Turbines on Birds, Proc. 1996 European Union Wind
Energy Conference, 156-159.
Colson, E. W. (1995) Avian Interactions with Wind Energy Facilities: A Summary, Proc. Windpower
'95. AWEA, 77-86.
Connors, S. R. (1996) Informing Decision Makers and Identifying Niche Opportunities for
Windpower: Use of Multiattribute Trade off Analysis to Evaluate Non-dispatchable Resources,
Energy Policy, 24(2), 165-176.
Dooling R (2002) Avian hearing and avoidance of wind turbines. NREL/TP-500-30844, National
Renewable Energy Laboratory, Colorado
Erickson WP, Strickland MD, Johnson GD, Kern JW (2000) Examples to statistical methods to assess
risks of impacts to birds from wind plants. Proceedings of National Avian-Wind Power Planning
meeting III, San Diego, California pp 172-182
European Commission (1998) Wind Energy - The Facts.
Fenhann J, Kilde NA, Runge E, Winther M, Boll Illerup J (1977) Inventory of emissions from Danish

| Environmental Aspects and Impacts Pavesi - 249 -


sources, 1972-1995. Riso National Environmental Research Institute
Germanisher Lloyd (1994) Regulation for the Certification of Wind Energy Conversion Systems,
Supplement to the 1993 Edition, Hamburg.
Gipe, P. (1 995) Wind Energy Comes of Age, Wiley, New York.
Goodman, N. (1997) The Environmental Impact of Windpower Development in Vermont: A Policy
Analysis, Proc, Windpower '97. AWEA, 299-308.
Green RH (1979) Sampling designs and statistical methods for environmental biologists. Wiley, New
York, NY
Howell JA (1997) Avian mortality at rotor swept area equivalents, Altamont Pass and Montezuma
Hills, California. Trans. Western Sector Wildlife Society, 33:24-29
http://www.greenontario.org
Hubbard, H. H., Shepherd, K. P. (1990) Wind Turbine Acoustics, NASA Technical Paper 3057
DOEA'AsAn0320-77.
Hunt WG (2000) A population study of Golden Eagles in Altamont Pass wind resource area:
Population trend analysis 1994-1997-Executive summary. Proceedings of National Avian-Wind
Power Planning meeting III, San Diego, California pp 15-17
IEA (1994) Expert group study on recommended practices for wind turbine testing and evaluation, 4.
acoustic noise emission from wind turbines. International Energy Agency
IEA (2003) Emissions. Key world energy statistics. International Energy Agency pp 44-46
IEC (2001) Wind turbine generator systems-Part 11 Acoustic noise measurement techniques. IEC
61400-11, No. 88/141/CDV, International Electrochemical commission, Geneva, Switzerland
Kenetech Wind power (1994) Kenetech Wind power avian research programme update. Kenetech
Wind power, Washington D.C.
Kidner, D. B. (1996) The Visual Impact of Taff Ely Wind Farm - A Case Study Using GIS. Wind
Energy Conversion 1996, BWEA, 205-21 1.
Klung H (2002) Noise from wind turbines-standards and noise reduction procedures. Forum Acusticum
2002, Sevilla, Spain
Lowther SM, Tyler S (1996) A review of impacts of wind turbines on birds in the UK. Report No.
W/13/00426/REP3, Energy Technology Support Unit (ETSU)
Marti Montes R, Barrios Jaque L (1995) Effects on wind turbine power plants on the avifauna in the
Campo de Gibraltar region. Spanish Ornithological Society.
Morrison ML (2000) The role of visual activity in bird-wind turbine interactions. Proceedings of
National Avian-Wind Power Planning meeting III, San Diego, California pp 28-30
MosArt Associates (2000) Landscape assessment for wind farm planning and design-character and
sensitivity. Final report, Altener project AL/98/542 pp 1-18
Neilson FB (1996) Wind turbine and the landscape: Architecture and aesthetics. Birk Nielsens
Tegnestue, Soendergade, Denmark
NIRS (1989) Committee finds massive sea life kills from San Onofre. Nuclear Information and
Resource Service Groundswell 11 (2&3)
NWCC (National Wind Coordinating Committee) (1998) Permitting of Wind Energy Facilities: A
Handbook, RESOLVE, Washington, DC.
NWCC (National Wind Coordinating Committee) (1999) Studying Wind Energy/bird Interactions: A
Guidance Document, RESOLVE, Washington, DC.
Orloff S, Flannery A (1992) A continued examination of avian mortality in the Altamont Pass wind
resource area. Report P700-96-004CN, Ibis Environmental Services and BioSystems Analysis

| Environmental Aspects and Impacts Pavesi - 250 -


Inc.
Orloff, S., Flannery, A. (1992) Wind Turbine Effects on Avian Activity, Habitat Use, and Mortality in
Altamont Pass and Solano County Wind Resource Areas: 1989-1991, California Energy
Commission Report No. P700-92-001.
Ratto, C. F., Solari, G. (1998) Wind Energy and Landscape, A. A. Balkema, Rotterdam.
Robotham, A. J. (1992) Progress in the Development of a Video Based Wind Farm Simulation
Technique. Wind Energy Conversion 1992, BWEA, 35 1-353.
Rogers AL, Manwell JF (2002) Wind turbine noise issues – A white paper. Renewable Energy
Research Laboratory, Centre for Energy Efficiency and Renewable Energy, Department of
Mechanical and Industrial Engineering, University of Massachusetts at Amherst, Amherst
Schleisner L (2000) Life cycle assessment of a wind farm and related externalities. Renewable Energy
20: 279-288
Schmidt E, Piaggio AJ, Bock CE, Armstrong DM (2003) National wind technology centre site
environmental assessment – Final report. NREL/SR-500-32981, National Renewable Energy
Laboratory, Colorado
Schwela D (1998) WHO guidelines on community noise. Proceedings of ICBEN 98, Sydney, pp 475-
480
Sengupta, D. L., Senior, T.B.A. (1994) Electromagnetic Interference from Wind Turbines, Chapter 9 in
Wind Turbine Technology, ed. D. A. Spera, ASME, New York.
Sengupta, D. L., Senior, T.B.A., Ferris, J. E. (1983) Study of Television Interference by Small Wind
Turbines: Final Report, Solar Energy Research Institute Report SERUSTR-215-1881, "IS.
Senior, T.B.A., Sengupta, D. L. (1983) Large Wind Turbine Handbook Television Interference
Assessment, Solar Energy Research Institute Report SERI/STR-215-1879, NTIS.
SEPA (2003) Noise annoyance from wind turbines-A review. Swedish Environmental Protection
Agency, Report 5308
Sinclair, K. C. (1999) Status of the U.S. Department of Energy/National Renewable Energy Laboratory
Avian Research Program, Proceedings Windpower '99, AWEA.
Sinclair, K. C., Momson, M. L. (1997) Overview of the U.S. Department of Energy/National
Renewable Energy Laboratory Avian Research Program, Proc. Windpower '97, AWEA, 273-279.
Stanton, C. (1994) The Visual Impact and Design of Wind Farms in the Landscape, Wind Energy
Conversion 1994, BWEA, 249-255.
Stanton, C. (1995) Wind Farm Visual Impact and its Assessment, Wind Directions, BWEA, August, 8-
9.
Stanton C (1996) The landscape impact and visual design of wind farms. ISBN 1-901278-00-X, Heriot-
Watt University, United Kingdom
Still, D., Painter, S., Lawrence, E. S., Little, B., Thomas, M. (1996) Birds, Wind Farms, and Blyth
Harbor, Wind Energy Conversion 1996, BWEA, 175-183.
Taylor, D. C., Durie, M. J. (1995) Wind Farm Visualisation, Wind Energy Conversion 1995, BWEA,
413-416.
Turner, D. M. (1989) An Analysis oE Blade Throw from Wind Turbines, in Wind Energy and the
Environment, ed. D. T. Swift-Hook, Peter Peregrinus, London.
Van Kats, P. J., Van Rees, J. (1989) Large Wind Turbines: A Source of Interference for TV Broadcast
Reception, Chapter 11 in Wind Energy and the Environment, ed. D. T. Swift-Book, Peter
Peregrinus, London.
Wagner, S., Bareib, R., Guidati, G. (1996) Wind Turbine Noise, Springer, Berlin.

| Environmental Aspects and Impacts Pavesi - 251 -


Wolf, B. (1995) Mitigating Avian Impacts: Applying the Wetlands Experience to Wind F m s , Proc.
Windpower '95, AWEA, 109-1 16.
Wolsink M, Sprengers M, Keuper A, Pedersen TH, Westra CA (1993) Annoyance from wind turbine
noise in sixteen sites in three countries. European community wind energy conference, Lubeck,
Travemunde pp 273- 276

| Environmental Aspects and Impacts Pavesi - 252 -

You might also like