Students Notes
Students Notes
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 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
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 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
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
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.
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
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.
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).
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
Figure 4.1.8 Altitude dependence of temperature, pressure and density for standard atmosphere.
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.
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).
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
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
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.
4.5.2 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)
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.
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.
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.
…………………………..
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
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
pu 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.
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 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
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.
ku
k 1
u k
f u exp (k>0, u>0, c>1) (5.6.17)
cc 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
.
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)
ku
k 1
u k
u exp du
0
cc c
If we make the change of variable
k
u
x (5.6.19)
x
then the mean wind speed can be written
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 y1e 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
2 1
1
k
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)
We may think of this value as the true or actual power in the wind if the probabilities p(ui) are
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.
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)
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
ln ln 1 F u k ln u k ln c (5.6.42)
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
ka
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
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
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
exp
1
f u (5.6.60)
2 22
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
uu
q
or (5.6.62)
u u q
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
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
x
j1
j
x
exp x
n
i j
j1
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
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
The air that passes through the disc undergoes an overall change in velocity, U - Uw and a rate of
p
d pd A d U U w A d U d
(6.1.4)
U U w A d U 1 a
p
d pd U2 U 2w
1
2
(6.1.9)
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 2Ad 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
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
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
P m U2 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
P 2Ad U3 a 1 a
2
(6.3.3)
Hence
2Ad U3 a 1 a 2A 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
The area of the ring is A d 2rr therefore the incremental shaft power is, from Equation
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
27
Which is precisely the same as for the non-rotating wake case.
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.
r (6.6.14)
1 a '
dM 1 8rU a 1 a
3 2
dr 2 1 a '
Power
dP dM
dr dr
1 8rU a 1 a
3 2
(6.6.15)
2 1 a '
1 4R 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 2a '
1
2
2
(6.6.17)
2a ' 1 a ' 2 r 2
The resulting thrust on an annular element, dF , is:
dF pd pd 2rdr
(6.6.18)
4a ' 1 a ' 2 r 3dr
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 4R 2 a ' 1 a 3rd r (6.6.20)
2
The power generated at each element, dP , is given by:
U3
dP 4R 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:
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 dr, 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 11 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 13a 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.
.
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
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.
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 rr (7.2.12)
2
The rate of change of angular momentum of the air passing through the annulus is
U 1 a r2a 'r2rr 4U r a ' 1 a r 2r (7.2.13)
Equating the two moments
1
W 2 n b c CL sin Cd cos rr 4U r a ' 1 a r 2r (7.2.14)
2
Simplifying,
W2 c
2
n b CL sin Cd cos 82a ' 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
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
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).
7.4.1 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
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
U2
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 8a 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
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.
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).
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
Figure 9.3.11 Power Curves for Different Negative Pitch Angles: 40 m Diameter Rotor
Rotating at 33 r.p.m.
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 VN 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.
9.5.4 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.
Figure 9.5.1 Passive Control of Tip Blade, Using Screw on Tip Shaft and Spring
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:
FV 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.
E IR T PV f V dV (10.4.3)
VI
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
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
kV V
k
E RO PR T e (10.4.10)
VR
c c
As
f V dV F V (10.4.11)
E RO Pr T e Xr e Xo (10.4.12)
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
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
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)
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.
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
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
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
a AV3 VI VI
2 2
1 a Vd3 Gd
QV Cpd pd T VDT (10.7.24)
8 w gh d Npd
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
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 Gd O 2 X
V
Table 13. 1 Stack emissions of coal, gas, and wind power plants (kg/MWh)
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
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 11.1.1 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
Figure 11.4.1 Eight categories of wind turbines in the Altamont pass (Orloff and Flannery,
1992)
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.
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.
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)
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 11.6.1 Wind turbine noise assessment factors (Hubbard and Shepherd, 1990)
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).
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).
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.
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
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).
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)
Class
Group Parameter Class 2
3
Turbine configuration Hub height X X
Type of tower (upwind or downwind) X
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.
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
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).
11.9.1 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