Skip to main content

Wind Turbine Power Performance and Application to Monitoring

  • Chapter
  • First Online:
Handbook of Wind Power Systems

Part of the book series: Energy Systems ((ENERGY))

Abstract

The concept of power performance is introduced as the ability of a wind turbine to extract power from the wind. The general performance estimates such as the power coefficient or the theoretical power curve are introduced in laminar conditions. Following Betz’ limit, an upper limit for the power available in the wind is derived, as well the main sources of energy loss. This laminar theory is too simple to describe operating wind turbines, and turbulent and atmospheric effects call for statistical tools. An IEC norm defines the international standard to measure and analyze power performance. The resulting IEC power curve gives a first estimation, and can be used to evaluate the annual energy production. An alternative is proposed with the Langevin power curve, which quantifies the high-frequency dynamics of a wind turbine power production to changing wind speeds. This brings further insight on the overall performance, and allows for applications such as performance monitoring or power modeling.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 169.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 169.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    An actuator disc is an infinitely thin disc through which the air can flow without resistance, as proposed by Froude and Rankine's momentum theory [12].

  2. 2.

    A stream-tube is defined here as the stream of air particles that interact with the wind turbine.

  3. 3.

    Following Froude-Rankine, it can be shown that this value is the optimal value.

  4. 4.

    Other wind turbine designs involve fixed rotational frequency (called fixed-speed wind turbines) or fixed pitch angle (called fixed-pitch wind turbines). A more detailed description on control strategies is given in [3].

  5. 5.

    Stall effects are obtained when the angle of attack of an airfoil exceeds a critical value, resulting in a sudden reduction in the lift force generated. A detailed study on airfoil lift effects can be found in [19].

  6. 6.

    Additional considerations such as mechanical loads or power stability are usually taken into account as well [3], but reach out of the scope of this chapter.

  7. 7.

    \( u_{cut - in} \) represents the minimum wind speed such that the wind turbine can extract power, typically in the order of \( 3{-}4\,\text{m/s} \).

  8. 8.

    \( u_{r} \) represents the rated wind speed at which the wind turbine extracts the rated, maximum allowed power \( P_{r} \), typically in the order of \( 12{-}15\,\text{m/s} \).

  9. 9.

    \( u_{cut - out} \) represents the maximum wind speed at which the wind turbine can safely extract power, typically in the order of \( 25{-}35\,\text{m/s} \).

  10. 10.

    Although no unique, clear definition of gust exists, one can see a wind gust as a rapid change of wind speed (and possibly direction). Gusts are extreme events.

  11. 11.

    The notion of intermittency is related to the probability of a process to sustain extreme events. It can be identified as a large deviation from the Gaussian distribution far away from the mean value. Extreme events such as gusts yield intermittent PDFs.

  12. 12.

    Additional correction of the measured data should be performed using temperature and pressure measurements.

  13. 13.

    To some extent the power output also fluctuates on short time scales, but its high-frequency dynamics are limited by the inertia of the wind turbine.

  14. 14.

    \( \sigma^{2} = \overline{{(u - \overline{u} )^{2} }} = \overline{{u^{\prime}(t)^{2} }} \).

  15. 15.

    Assuming a cubic power curve \( P(u) \propto u^{3} \), \( P(u) \) has non-zero derivatives up to 3rd-order. Moreover, the transition point to rated power may have non-zero derivatives of arbitrary order, see Fig. 12.

  16. 16.

    The subscript L stands for ``Langevin'' as \( P_{L} (u) \) will be associated to the formalism of the Langevin equation.

  17. 17.

    In former publications on the topic, the Langevin power curve was called dynamical power curve or Markovian power curve. It is nonetheless the same approach.

  18. 18.

    This equation is the reason for the name of the Langevin power curve.

  19. 19.

    \( D^{(1)} \) and \( D^{(2)} \) are the first two Kramers-Moyal coefficients.

  20. 20.

    To separate stable (attractive) from unstable (repulsive) fixed points, also the slope of \( D^{(1)} (P) \) must be considered.

  21. 21.

    Assuming that the measurement period is sufficiently long to reach statistical convergence.

  22. 22.

    Another alternative is given by ultrasonic anemometers which can estimate at once the wind speed and direction.

  23. 23.

    For reminder, the IEC method focuses on time scales of 10 min, while the Langevin approach investigates the dynamics in the order of few seconds.

  24. 24.

    Local in wind speed and power output.

  25. 25.

    Typical hub heights of commercial multi-MW class wind turbines are in the order of 100 m, justifying the interest for a portable LIDAR sensor, see Sect. 3.4.

  26. 26.

    A measurement of wind speed over 1 year covers the various wind situations resulting from various seasonal behaviors.

  27. 27.

    One should note here that \( \lambda \) is not the tip speed ratio of a wind turbine, but a parameter of the Weibull distribution.

  28. 28.

    The IEC norm [10] refers to the Rayleigh distribution, which is a special case of the Weibull distribution for \( k = 2 \).

  29. 29.

    An over-sensitive procedure might indicate non-existing anomalies, while an under-sensitive procedure would fail to detect a major malfunction.

  30. 30.

    The reference time is chosen when the wind turbine is believed to work with full capacity.

  31. 31.

    This effective area is the one that enters the formula to calculate drag and lift forces.

  32. 32.

    For the Betz optimum, \( u_{2} = 2u_{1/3} \).

  33. 33.

    The force in the axial direction does not contribute to the power production of a wind turbine but to the thrust on it.

  34. 34.

    The Langevin equation relates directly the incoming wind speed and the power output. Many other variables are involved in intermediate steps of the conversion, which should be modeled by a set of mutli-dimensional deterministic differential equations. All these degrees of freedom are modeled by the one-dimensional stochastic Langevin equation instead.

  35. 35.

    A constant diffusion function yields additive noise. More complex systems such as turbulence-driven systems display multiplicative noise and a non-constant diffusion function.

  36. 36.

    A realistic model integrates a finite reaction time due to the inertia of the wind turbine to changing wind speeds.

  37. 37.

    An initial condition \( P(t = 0) \) for the power output is also necessary. However, the result depends only poorly on this value, as the dynamics will adjust rapidly to the given wind speed.

References

  1. Anahua E, Barth S, Peinke J (2008) Markovian power curves for wind turbines. Wind Energy 11(3):219–232

    Article  Google Scholar 

  2. Betz A (1927) Die windmühlen im lichte neuerer forschung. Die Naturwissenschaften 15:46

    Article  Google Scholar 

  3. Bianchi FD, De Battista H, Mantz RJ (2006) Wind turbine control systems, 2nd edn. Springer, Berlin

    Google Scholar 

  4. Böttcher F, Peinke J, Kleinhans D, Friedrich R (2007) Handling systems driven by different noise sources—implications for power estimations. In: Wind energy. Springer, Berlin, pp 179–182

    Google Scholar 

  5. Böttcher F, Barth S, Peinke J (2007) Small and large fluctuations in atmospheric wind speeds. Stoch Environ Res Ris Assess 21:299–308

    Article  Google Scholar 

  6. Burton T, Sharpe D, Jenkins N, Bossanyi E (2001) Wind energy handbook. Wiley, New York

    Book  Google Scholar 

  7. Gottschall J (2009) Modelling the variability of complex systems by means of Langevin processes. PhD thesis, Carl von Ossietzky Universität Oldenburg, Germany

    Google Scholar 

  8. Gottschall J, Peinke J (2007) Stochastic modelling of a wind turbine’s power output with special respect to turbulent dynamics. J Phys: Conf Ser 75:012045

    Google Scholar 

  9. Gottschall J, Peinke J (2008) How to improve the estimation of power curves for wind turbines. Environ Res Lett 3(1):015005 (7 pp)

    Article  Google Scholar 

  10. IEC (2005) Wind turbine generator systems, part 12: wind turbine power performance testing. International Standard 61400-12-1, International Electrotechnical Commission

    Google Scholar 

  11. Milan P, Mücke T, Wächter M, Peinke J (2010) Two numerical modeling approaches for wind energy converters. In: Proceedings of computational wind engineering 2010

    Google Scholar 

  12. Rankine WJM (1865) On the mechanical principles of the action of propellers. Trans Inst Naval Architects 6:13–39

    Google Scholar 

  13. Rauh A, Seelert W (1984) The Betz optimum efficiency for windmills. Appl Energy 17:15–23

    Article  Google Scholar 

  14. Rauh A (2008) On the relevance of basic hydrodynamics to wind energy technology. Nonlinear Phenom Complex Syst 11(2):158–163

    MathSciNet  Google Scholar 

  15. Rauh A, Peinke J (2004) A phenomenological model for the dynamic response of wind turbines to turbulent wind. J Wind Eng Ind Aerodyn 92(2):159–183

    Article  Google Scholar 

  16. Richardson LF (1922) Weather prediction by numerical process. Cambridge University Press, Cambridge

    MATH  Google Scholar 

  17. Risken H (1984) The Fokker-Planck equation. Springer, Berlin

    Book  MATH  Google Scholar 

  18. Rosen A, Sheinman Y (1994) The average power output of a wind turbine in turbulent wind. J Wind Eng Ind Aerodyn 51:287

    Article  Google Scholar 

  19. Schneemann J, Knebel P, Milan P, Peinke J (2010) Lift measurements in unsteady flow conditions. Proceedings of the European Wind Energy Conference 2010

    Google Scholar 

  20. Twele J, Gasch R (2005) Windkraftanlagen. Teubner B.G. GmbH

    Google Scholar 

  21. Wächter M, Gottschall J, Rettenmeier A, Peinke J (2008) Dynamical power curve estimation using different anemometer types. In: Proceedings of DEWEK, Bremen, 26–27 Dec 2008

    Google Scholar 

Download references

Acknowledgments

Parts of this work have been financially supported by the German Ministry for Environment (BMU) under grant number 0327642A. The authors would like to thank Stephan Barth, Julia Gottschall, Edgar Anahua and Michael Hölling for their pioneering work on the Langevin approach, as well as for stimulating discussions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Patrick Milan .

Editor information

Editors and Affiliations

Appendices

Appendix 1: Aerodynamics of Rotor Blades

The essential (mechanical) element of a wind turbine is the rotor, that transforms the power of the wind into a rotational or mechanical power. The ideal requirements are:

  • the rotation should be steady and smooth;

  • dynamical loads should be minimal;

  • the regulation should be done without sudden jumps.

The number of blades, their profile and design should guarantee these features. Modern wind turbines rotate due to the lift forces \( F_{L} \) acting on the airfoils. For an airfoil the effective areaFootnote 31 can be expressed in terms of the depth \( t \) and the wingspan \( b \) (normally equal to the rotor radius \( R \)), such that

$$ \begin{gathered} F_{D} = C_{D} (\alpha )\frac{1}{2}\rho c^{2} (t \cdot b) \hfill \\ F_{L} = C_{L} (\alpha )\frac{1}{2}\rho c^{2} (t \cdot b) , \hfill \\ \end{gathered} $$
(31)

where \( \alpha \) is the angle of attack, as displayed in Fig. 21. The lift-drag ratio \( F_{L} /F_{D} \) relates to the quality of the airfoil. The larger the ratio, the better the quality.

Fig. 21
figure 21

Cut through an airfoil to illustrate the forces acting on it. The depth \( t \) is given by the distance between the leading and the trailing edge. The wingspan \( b \) is the length of the airfoil, here perpendicular to the illustrated plane

In Fig. 21, the velocity vector \( \mathbf{c} \) gives the wind velocity in the frame of reference of the airfoil. The wind velocity is \( \mathbf{u}_{2} \) in the frame of the ground, but the rotational motion of the rotor must be considered for the motion of the wind with respect to the blades. Hence, \( \mathbf{c} \) is the superposition of the horizontal axial velocity \( \mathbf{u}_{2} \) Footnote 32 and of the rotational velocity \( v = \omega r \), such that

$$ c^{2} (r) = \left( {2u_{1} /3} \right)^{2} + \left( {\omega r} \right)^{2} . $$
(32)

The rotor feels the effective wind speed \( \mathbf{c} \). This is illustrated in Fig. 22.

Fig. 22
figure 22

Cut through a rotating airfoil. The rotational velocity \( \omega r \) is perpendicular to the axial velocity vector \( {\mathbf{u}}_{2} \). \( \beta \) denotes the angle between the resulting velocity \( \mathbf{c} \) and the rotational direction

Instead of integrating the lift and drag forces on the entire airfoil, one can estimate the local force on each infinitesimal element. Also, the total force is divided into its rotational component \( F_{r} \) and its axial component \( F_{a} \). Considering a cut \( dr \) at \( r \) in the polar plane of the rotor, the resulting force is

$$ \begin{aligned} F_{r} & = \frac{\rho }{2}c^{2} \cdot t \cdot dr\left[ {C_{L} \cos(\beta ) - C_{D} \sin(\beta )} \right] \hfill \\ F_{a} & = \frac{\rho }{2}c^{2} \cdot t \cdot dr\left[ {C_{L} \sin(\beta ) + C_{D} \cos(\beta )} \right]. \hfill \\ \end{aligned} $$
(33)

Also, on can estimate

$$ \tan(\beta ) = \frac{\omega r}{{u_{2} }} = \frac{\omega R}{{u_{1} }}\frac{r}{R}\frac{{u_{1} }}{{u_{2} }} = \frac{3}{2}\lambda \frac{r}{R} . $$
(34)

The idea is to construct the blades in such a way that for each infinitesimal radial annulus they extract the infinitesimal (Betz) optimal power out of the wind

$$ dP_{r,Betz} = \frac{16}{27} \cdot \frac{\rho }{2} \cdot u_{1}^{3} \cdot (2\pi rdr) . $$
(35)

This power also reads

$$ dP = z \times F_{r} \times \omega r , $$
(36)

where \( z \) denotes the number of blades, \( \omega r \) the velocity in rotational direction and \( F_{r} \) the force in this direction.Footnote 33 Inserting Eq. (33) and combining Eqs. (35) and (36), the optimal value of the depth \( t \) as a function of \( r \) can be determined. Assuming that \( C_{D} { \ll }C_{L} \) and a sufficiently large tip speed ratio (for details see [20]), the profile of the airfoil \( t(r) \) reads

$$ \begin{aligned} t(r) & \approx \frac{16\pi }{9}\frac{{R^{2} }}{{zC_{L} r\lambda^{2} }} \hfill \\ & \propto z^{ - 1} \cdot C_{L}^{ - 1} \cdot r^{ - 1} \cdot \lambda^{ - 2} . \hfill \\ \end{aligned} $$
(37)

This has an important consequence on the design of rotor blades. The depth decreases with increasing number of blades, larger lift coefficient, increasing radius and especially increasing tip speed ratios. This explains why fast rotating wind turbines tend to have only two or three narrow blades while old western-mill machines have many, rather broad blades.

Appendix 2: A Relaxation Model for the Power Output

As introduced in Sect. 3.3.2, the power output of wind turbines is assumed to be solution of a Langevin equation

$$ \frac{d}{dt}P(t) = D^{(1)} (P) + \sqrt {D^{(2)} (P)} \times \varGamma (t) . $$
(38)

For reminder, the power value \( P \) and the functions \( D^{(1)} \) and \( D^{(2)} \) are conditioned on the wind speed bins, as done in the main text. The subscript \( _{i} \) indicating the wind speed bin was dropped for simplicity. \( D^{(1)} \) represents the deterministic dynamics of the conversion process, that always push the power output towards the Langevin power curve \( P_{L} (u) \). Additional random fluctuations are superposed as a simplified model for all the microscopic degrees of freedom acting on the conversion process.Footnote 34 A simple but rather realistic ansatz for \( D^{(1)} \) would be

$$ D^{(1)} (P) = \alpha \left(P_{theo} (u(t)) - P(t)\right) , $$
(39)

where \( D^{(1)} \) linearly drives the power output towards the instantaneous value of the theoretical power curve \( P_{theo} (u(t)) \), which might read

$$ P_{theo} (u) = \left( {\begin{array}{ll} {P_{r} \left( {\frac{u}{{u_{r} }}} \right)^{3} } &{{\text{for}}\;{\text{u}} \le{\text{u}}_{\text{r}} ,} \\ {P_{r} } & {{\text{for}}\;{\text{u}} \ge {\text{u}}_{\text{r}} .} \\ \end{array} } \right) $$
(40)

Assuming Eq. (39) and a constant diffusion function \( D^{(2)} (P) = \beta \),Footnote 35 the Langevin equation becomes a relaxation model for the power output

$$ \frac{d}{dt}P(t) = \alpha \left( {P_{theo} (u(t)) - P(t)} \right) + \sqrt \beta \times \varGamma (t) . $$
(41)

Equation (41) is a phenomenological model for the power signal. This special case of the Langevin process is mathematically called an Ornstein–Uhlenbeck process [17].

Equation (41) is a simplified model for the power output, where the wind turbine design is described through the parameters \( \alpha \) and \( \beta \), as well as the power curve \( P_{theo} (u) \). The parameter \( \alpha \) is related to the reaction time of the model wind turbine,Footnote 36 while \( \beta \) quantifies the strength of the stochastic noise. \( \varGamma (t) \) is a Gaussian-distributed white noise with mean value 0 and variance 2, which can be generated easily from most mathematical softwares. Using a wind speed time series \( u(t) \) as an input for the model equation,Footnote 37 a time series of power output \( P(t) \) can be generated at the same sampling frequency. An example is provided in Fig. 23.

Fig. 23
figure 23

a Power signal measured (red line). b Power signal modeled (blue line) following Eq. (41) with \( \alpha = 0.005\, / \text{s} \), \( \beta = 0.5\, {\text{W/s}}^{2} \), \( u_{r} = 13\,\text{m/s} \) and \( P_{r} = 1\, \text{W} \). The wind speed signal used was measured simultaneously as the measured power output

From Fig. 23, it can be seen that the relaxation model manages to estimate the power output of a wind turbine to a first approximation. Fluctuations and their statistics are more difficult to reproduce than long-time behavior, which is mostly driven by the changes in wind speed rather than by the stochastic fluctuations. More advanced methods are being developed, as introduced in [11], where \( D^{(1)} \) and \( D^{(2)} \) are not assumed but estimated from measurement data.

Rights and permissions

Reprints and permissions

Copyright information

© 2013 Springer-Verlag Berlin Heidelberg

About this chapter

Cite this chapter

Milan, P., Wächter, M., Peinke, J. (2013). Wind Turbine Power Performance and Application to Monitoring. In: Pardalos, P., Rebennack, S., Pereira, M., Iliadis, N., Pappu, V. (eds) Handbook of Wind Power Systems. Energy Systems. Springer, Berlin, Heidelberg. https://doi.org/10.1007/978-3-642-41080-2_20

Download citation

  • DOI: https://doi.org/10.1007/978-3-642-41080-2_20

  • Published:

  • Publisher Name: Springer, Berlin, Heidelberg

  • Print ISBN: 978-3-642-41079-6

  • Online ISBN: 978-3-642-41080-2

  • eBook Packages: EnergyEnergy (R0)

Publish with us

Policies and ethics