control of variable-speed wind turbines

Upload: rene-alexander-barrera-cardenas

Post on 06-Apr-2018

224 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/3/2019 Control of Variable-speed Wind Turbines

    1/12

    70 IEEE CONTROL SYSTEMS MAGAZINE JUNE 2006 1066-033X/06/$20.002006IEEE

    Wind energy is the fastest-growing energy

    source in the world, with worldwide

    wind-generation capacity tripling in the

    five years leading up to 2004 [1].

    Because wind turbines are large, flexible

    structures operating in noisy environments, they pre-

    sent a myriad of control problems that, if solved, could

    reduce the cost of wind energy. In contrast to constant-

    speed turbines (see Wind Turbine Development and

    Types of Turbines), variable-speed wind turbines are

    designed to follow wind-speed variations in lowwinds to maximize aerodynamic efficiency. Standard

    control laws [2] require that complex aerodynamic

    properties be well known so that the variable-speed

    turbine can maximize energy capture; in practice,

    uncertainties limit the efficient energy capture of a

    variable-speed turbine. The turbine used as a model

    for this articles research is the Controls Advanced

    Research Turbine (CART) pictured in Figure 1. CART

    is located in Golden, Colorado, at the U.S. National

    Renewable Energy Laboratorys National Wind Tech-

    nology Center (see The National Renewable Energy

    Laboratory and National

    Wind Technology Center).

    A modern utility-scale

    wind turbine, as shown in

    Figure 2, has several levels of control systems. On the

    uppermost level, a supervisory controller monitors

    the turbine and wind resource to determine when

    the wind speed is sufficient to start up the turbine

    and when, due to high winds, the turbine must be

    shut down for safety. This type of control is the dis-

    crete if-then variety. On the middle level is turbinecontrol, which includes generator torque control,

    blade pitch control, and yaw control. Generator

    torque control, performed using the power electron-

    ics, determines how much torque is extracted from

    the turbine, specifically, the high-speed shaft. The

    extracted torque opposes the aerodynamic torque

    provided by the wind and, thus, indirectly regulates

    the turbine speed. Depending on the pitch actuators

    and type of generator and power electronics, blade

    pitch control and generator torque control can oper-

    ate quickly relative to the rotor-speed time constant.

    STANDARD AND ADAPTIVE TECHNIQUES

    FOR MAXIMIZING ENERGY CAPTURE

    KATHRYN E. JOHNSON, LUCY Y. PAO, MARK J. BALAS, and LEE J. FINGERSH

    NATIONAL RENEWABLE ENERGY LABORATORY

  • 8/3/2019 Control of Variable-speed Wind Turbines

    2/12

    Yaw control, which rotates the nacelle to point into the

    wind, is slower than generator torque control and blade

    pitch control. Due to its slowness, yaw control is of less

    interest to control engineers than generator torque control

    and blade pitch angle control.

    On the lowest control level are the internal generator,

    power electronics, and pitch actuator controllers, which

    operate at higher rates than the turbine-level control. These

    low-level controllers operate as black boxes from the per-

    spective of the turbine-level control. For example, the gener-

    ator and power electronics controllers regulate the generator

    and power electronics variables to achieve the desired gen-

    erator torque, as determined by the turbine-level control.

    The low-level controllers depend on the types of generator

    and power electronics, but the turbine-level control does

    not. For example, CART has a squirrel-cage induction gen-

    erator and full-processing pulse-width modulation power

    electronics. If the generator torque controller controls the

    high-speed shaft torque, then the stability analysis of the

    turbine-level control does not depend on these details. In

    A Rotor swept area (m2)

    Cp Rotor power coefficient (dimensionless)

    Cpmax Maximum rotor power coefficient (dimensionless)

    Cq Rotor torque coefficient (dimensionless)

    J Rotor inertia (kg-m2)

    K Standard torque control gain (kg-m2)

    M Adaptive torque control gain (m5)

    M+ Simulation-derived prediction of optimal torque

    control gain (m5)

    M Turbines true optimal torque control gain (possiblyunknown) (m5)

    P Turbine (rotor) power (kW)

    P0 Symmetric quadratic curve coefficient (dimensionless)

    Pcap Captured power (kW)

    Pfavg Average captured power divided by average wind

    power over a given time period (dimensionless)

    Pwind Power available in the wind (kW)

    Pwy Power available in the wind, with approximate yaw

    error factor included (kW)

    R Rotor radius (m)

    a Symmetric quadratic curve coefficient (m10)

    b Damping coefficient (kg-m2/s)

    fs Sampling frequency (Hz)

    k Adaptive controllers discrete-time index

    n Number of steps in adaptation period

    v Wind speed (m/s)

    Blade pitch angle (deg)M Positive gain in gain adaptation law (m

    5)

    Tip-speed ratio (TSR) (dimensionless)

    TSR corresponding to Cpmax (dimensionless)

    Air density (kg/m3)

    aero Aerodynamic torque (N-m)

    c Generator (control) torque (N-m)

    Yaw error (deg)

    Rotor angular speed (rad/s)

    JUNE 2006 IEEE CONTROL SYSTEMS MAGAZINE 71

    Wind-powered machines have been used by humans for cen-

    turies. Most familiar are the historical many-bladed windmills

    used for milling grain, the earliest versions of which appeared

    during the 12th century [21]. Water-pumping wind machines

    appeared in the United States in the mid-19th century, while themodern era of wind turbine generators began in the 1970s [21].

    These modern horizontal-axis wind turbines typically have two or

    three blades and can be either upwind (with the rotor spinning on

    the upwind side of the tower) or downwind. Horizontal-axis wind

    turbines range in size from small home-based turbines of a few

    hundred watts to utility-scale turbines up to several megawatts.

    Most modern utility-scale turbines operate in variable-speed

    mode with the turbine speed changing continuously in response

    to wind gusts and lulls. Although costly power electronics are

    required to convert the variable-frequency power to the fixed utili-

    ty grid frequency, variable-speed turbines can spend more time

    operating at maximum aerodynamic efficiency than constant-

    speed turbines. In addition, variable-speed turbines often endure

    smaller power fluctuations and operating loads than constant-

    speed turbines. Constant-speed turbines are connected directly

    to the utility grid, which eliminates the requirement for power elec-

    tronics. A constant-speed machines fixed generator frequencyforces the turbines mechanical components to absorb much of

    the increased energy of a wind gust until the turbines power reg-

    ulation system can respond. On a variable-speed machine, how-

    ever, the rotor speed can increase, absorbing a great deal of

    energy due to the large rotational inertia of the rotor.

    For modern turbines and power electronics systems, the

    increased efficiency and lower loads of variable-speed turbines pro-

    vide enough benefit to make the power electronics cost effective.

    The wind industry trend is thus to design and build variable-speed

    turbines for utility-scale installations. Controlling these modern tur-

    bines to minimize the cost of wind energy is a complex task, and

    much research remains to be done to improve the controllers.

    Wind Turbine Development and Types of Turbines

    Nomenclature

  • 8/3/2019 Control of Variable-speed Wind Turbines

    3/12

    this project, we ignore the particulars of the high- and low-

    level controls and focus on the turbine-level control.

    Variable-speed wind turbines have three main regions of

    operation. A stopped turbine or a turbine that is just starting

    up is considered to be operating in region 1. Region 2 is an

    operational mode with the objective of maximizing wind

    energy capture. In region 3, which encompasses high wind

    speeds, the turbine must limit the captured wind power so

    that safe electrical and mechanical loads are not exceeded.

    For each region, the solid curve in Figure 3 illustrates the

    desired power-versus-wind-speed relationship for a vari-

    able-speed wind turbine with a 43.3-m rotor diameter.

    In Figure 3, the power coefficient Cp is defined as the

    ratio of the aerodynamic rotor power P to the power Pwindavailable from the wind, that is,

    Cp =P

    Pwind. (1)

    The available power Pwind is given by

    Pwind =12 Av

    3, (2)

    where is the air density,A is the rotor swept area, and v is

    the wind speed. The aerodynamic rotor power is given by

    P= aero, (3)

    where aero is the aerodynamic

    torque applied to the rotor by the

    wind and is the rotor angular

    speed. In Figure 3, the dotted wind

    power curve represents the power

    in the unimpeded wind passingthrough the rotor swept area,

    whereas the solid curve represents

    the power extracted by a typical

    variable-speed turbine. Because

    the wind can change speed more

    quickly than the turbine, there

    does not exist a static relationship

    between wind speed and turbine

    power in dynamic conditions.

    However, under steady-state con-

    ditions, a static relationship exists;

    the turbine power curve plotted in

    Figure 3 represents the power ver-sus wind speed relationship for a

    turbine withCp = 0.4.

    Classical techniques such as pro-

    portional, integral, and derivative

    (PID) control of blade pitch [3] are

    typically used to limit power and

    speed on both the low-speed shaft

    and high-speed shaft for turbines

    operating in region 3, while

    72 IEEE CONTROL SYSTEMS MAGAZINE JUNE 2006

    FIGURE 2 Major components of an upwind turbine, in which the wind hits the rotor before the

    tower. Unlike CART, this turbine rotor has three blades. Most turbines have a fixed-ratio gearbox,

    as shown, rather than a transmission, since it is not economical to build a transmission capable

    of withstanding a wind turbines high torques and extensive operating hours. The power electron-

    ics for a variable-speed turbine are usually located at the base of the tower. (Drawing courtesy of

    the U.S. Department of Energy.)

    Pitch

    WindDirection

    Low-SpeedShaft

    Rotor

    Brake

    Gear Box

    Yaw Drive

    Yaw Motor

    Blades Tower

    High-SpeedShaft

    Nacelle

    Wind Vane

    Controller

    GeneratorAnemometer

    FIGURE 1 CART at the National Wind Technology Center. CART is

    a 600-kW turbine with a 43.3-m rotor diameter used in advanced

    control experiments. The aim of these control experiments is to

    reduce the cost of wind energy, either by increasing the amount of

    energy extracted from the wind or by decreasing the turbines cost

    by reducing the stress on its components.

  • 8/3/2019 Control of Variable-speed Wind Turbines

    4/12

    generator torque control [4] is usually used in region 2. In [5],

    disturbance accommodating control is used to limit power and

    speed in region 3. The reduction of mechanical loads on the

    tower and blades is another area of turbine control research

    [6][8]. Finally, [9][12] use adaptive control to compensate for

    unknown and time-varying parameters in regions 2 and 3.

    Although specific techniques for controlling modern turbines

    are usually proprietary, we believe that only recently have tur-

    bine manufacturers begun to incorporate more modern and

    advanced control methods in commercial turbines. In part, the

    gap between the research and commercial turbine communi-

    ties is a result of the fact that few theoretically advanced con-

    trollers have been successfully tested on real turbines.

    In this article, we analyze the stability of a control sys-

    tem that has been tested on CART, focusing on adaptive

    generator torque control with constant blade pitch to maxi-

    mize energy capture of a variable-speed wind turbine

    operating in region 2. In [2], an adaptive strategy is shown

    to improve wind turbine performance. The focus of this

    article is stability analysis of the adaptive generator torquecontroller. We begin with a review of nonadaptive con-

    trollers, continue with a discussion of the adaptive con-

    troller of [2], and then proceed to the stability analysis.

    STANDARD VARIABLE-SPEED CONTROL LAW

    For variable-speed wind turbines operating in region 2, the

    control objective is to maximize energy capture by operat-

    ing the turbine at the peak of the Cp-TSR-pitch surface of

    the rotor, shown in Figure 4. The power coefficientCp(, )

    is a function of the tip-speed ratio (TSR) and the blade

    pitch . The TSR is defined as

    =R

    v . (4)

    Since, by (1), rotor power P increases with Cp, operation at

    the maximum power coefficient Cpmax is desirable. We note

    that Cp can be negative, which corresponds to operating

    the generator in reverse as a motor while drawing power

    from the utility grid. Also, the Cp surface changes when

    the condition of the blade surface changes. For example,

    icing or residue buildup on the blade typically shifts the Cpsurface downward, reducing energy capture. In this sec-

    tion, we assume the blades are clean.

    Figure 4 is based on the modeling software PROP [13],

    which uses blade-element momentum theory [14]. The

    PROP simulation was performed to estimate Cp for the600-kW two-bladed, upwind CART. Unfortunately, mod-

    eling tools such as PROP are of questionable accuracy; in

    fact, an NREL study [15] comparing wind turbine model-

    ing codes reports large discrepancies and an unknown

    level of uncertainty. Therefore, computer models are unre-

    liable for fixed-gain controller synthesis.

    A control law, which we refer to as the standard control,

    for region 2 operation of variable-speed turbines is to let the

    control torque c (that is, the generator torque) be given by

    JUNE 2006 IEEE CONTROL SYSTEMS MAGAZINE 73

    The National Renewable Energy Laboratoryand National Wind Technology Center

    The National Renewable Energy Laboratory (NREL) is a

    part of the U.S. Department of Energy (DOE) Office of

    Energy Efficiency and Renewable Energy. Located in Gold-

    en, Colorado, the laboratory began operating in 1977 as

    the Solar Energy Research Institute (SERI) and attained

    the national laboratory classification in 1991 when SERI

    was renamed NREL. NRELs mission statement summa-

    rizes the laboratorys research: NREL develops renewable

    energy and energy efficiency technologies and practices,

    advances related science and engineering, and transfers

    knowledge and innovations to address the nations energy

    and environmental goals.

    The National Wind Technology Center (NWTC) supports

    the U.S. wind industry by performing applied research and

    testing in conjunction with its industry partners. These indus-

    try partners range from large commercial turbine manufactur-

    ers to small distributed wind system developers, all of whomshare the goal of reducing the cost of wind energy. The

    NWTCs facilities include numerous turbine test pads, which

    currently test turbines ranging from 300 W to 600 kW; a

    dynamometer facility for testing advanced drive trains; an

    industrial user facility for testing new blade designs; a hybrid

    test facility, which allows testing of energy systems consist-

    ing of wind combined with solar, diesel, or other electricity

    sources; and two advanced research turbines. Together with

    NWTCs wind industry partners, researchers at the NWTC

    have helped to bring the cost of large-scale wind energy

    down from about US$0.80/kW-h in 1980 (todays dollars) to

    US$0.04US$0.06/kW-h today.

    FIGURE 3 Illustrative steady-state power curves. A variable-speed

    turbine attempts to maximize energy capture while operating in

    region 2. In region 3, the power is limited to ensure that safe electri-

    cal and mechanical loads are not exceeded.

    2,000

    1,800

    1,600

    1,400

    1,200

    1,000

    800

    600

    400

    200

    00 5 10

    Wind Speed (m/s)

    Power(kW)

    Wind PowerCp= 1

    15 20 25

    Region 1

    Region 2

    Region 3

    Turbine Power

    HighWind

    Cutout

    Cp= 0.4

  • 8/3/2019 Control of Variable-speed Wind Turbines

    5/12

    c = K2, (5)

    where the gain K is given by

    K =1

    2AR3

    Cpmax

    3, (6)

    R is the rotor radius, and is the tip-speed ratio at which

    the maximum power coefficientCpmax occurs.

    Next, assuming that the rotor is rigid, the angular accel-

    eration is given by

    =1

    J(aero c), (7)

    where J is the combined rotational inertia of the rotor,

    gearbox, generator, and shafts and the aerodynamic torque

    aero , derived from (1)(4), is given by

    aero =1

    2ARCq(,)v

    2, (8)

    where

    Cq(,) =Cp(,)

    (9)

    is the rotor torque coefficient. Since CART has a fairly rigid

    rotor, the rigid body model (7) is a valid approximation for

    the rotor dynamics.Now, substituting (8) and (5) into (7) and using (9) and

    (4) yields

    =1

    2JAR32

    Cp(,)

    3

    Cpmax

    3

    . (10)

    Since the rotor inertia J, the air density , the rotor swept

    area A, the rotor radius R, and the squared rotor speed 2

    are nonnegative, the sign of the angular acceleration

    depends on the sign of the difference in (10). When the tip-

    speed ratio > , it follows from (10) and the fact that

    Cp Cpmax that is negative and the rotor decelerates

    toward = . On the other hand, when < and

    Cp >Cpmax

    33, (11)

    it follows that is positive. The curve

    F() =Cpmax

    33

    is plotted as the dotted line in Figure 5, and CARTs PROP-

    derived Cp curve for a fixed pitch of 1 is the solid

    line. A pitch angle of 0 means that the blade chord line

    is approximately parallel to the rotor plane, although theexact angle depends on the amount of twist of the blade

    and the distance between the blade root and the chord line

    where the pitch angle is measured. The solid line in Figure 5

    is a two-dimensional slice of Figure 4. The inequality (11)

    is satisfied for tip-speed ratios ranging from about 3.3 to

    7.5. Thus, as long as CART has a tip-speed ratio of at least

    3.3, the standard control law (5) causes the speed of a well-

    characterized turbine to approach the optimal tip-speed

    ratio. Although easier to understand under constant wind

    74 IEEE CONTROL SYSTEMS MAGAZINE JUNE 2006

    FIGURE 4 Cp versus tip-speed ratio and pitch for CART. Since tur-

    bine power is proportional to the power coefficient Cp, the turbine is

    ideally operated at the peak of the surface. Blade pitch angle is a

    control variable, whereas tip-speed ratio is controlled indirectly using

    generator torque control. A turbines Cp surface can change due to

    icing, blade erosion, and residue buildup. Negative Cp corresponds

    to motoring operation during which the turbine draws energy from

    the utility grid.

    0.5

    0.4

    0.3

    0.2

    PowerCoefficientCp

    0.1

    1

    13

    15 1

    17

    3 1

    535

    79

    11

    0.1

    0.2

    0.3

    0.4

    0.5

    0.0

    Tip-Speed Ratio Pitch (deg)

    FIGURE 5 CARTs power coefficient Cp versus tip-speed ratio and

    cubic function F. The intersection of the solid and dotted lines at

    the tip-speed ratio = 7.5 indicates the optimal operating point in

    terms of energy capture. The cubic function F is derived from the

    standard control law, and the intersection points of the cubic func-

    tion and Cp curve are equilibrium points of the turbine operation.

    Theorem 2 shows that the equilibrium point = 7.5 is locally

    asymptotically stable.

    2 4

    0.4

    0.3

    0.2

    0.1

    0.0 6 8

    PowerCoefficientCp

    10 12 14

    F()

    Tip-Speed Ratio

  • 8/3/2019 Control of Variable-speed Wind Turbines

    6/12

    conditions, this behavior occurs in an averaged sense

    under time-varying wind conditions. We refer to the gain

    K corresponding to optimum tip-speed ratio operation as

    the optimal K.

    When the tip-speed ratio < 3.3, the inequality (11) is

    no longer satisfied, and the angular acceleration is nega-

    tive. In this case, the rotor speed slows toward zero.

    However, most turbines have separate control mechanisms

    to ensure that a low tip-speed ratio < 3.3 does not drive

    the rotor speed to zero when the wind speed is adequate

    for energy production. This article is concerned only with

    the torque control and, hence, does not consider these sep-

    arate control mechanisms. While the critical tip-speed

    ratios and control mechanisms are different for different

    turbines, the dynamics presented here approximate all

    variable-speed turbines using the standard control law (5).

    The above discussion assumes that the turbines prop-

    erties used to calculate the gain K in (6) are accurate, which

    is rarely the case. Also, over time, debris buildup and

    blade erosion change the Cp surface and thus Cpmax , withthe same effect as a suboptimally chosen K. The sensitivity

    of energy loss to errors in and the maximum power

    coefficient Cpmax is considered in [4], which concludes that

    a 5% error in the optimal tip-speed ratio can cause a sig-

    nificant energy loss of 13% in region 2. If the United

    States meets the American Wind Energy Associations goal

    of 100,000 MW of installed wind capacity by 2020, a 3%

    loss in total energy would equal US$300 million per year.

    The potential for cost savings motivates the development

    and investigation of an adaptive control approach that can

    improve energy capture.

    ADAPTIVE CONTROLFor region 2 operation, we now consider the adaptive con-

    troller [2] given by

    c =

    0, < 0,

    M2, 0,(12)

    where the adaptive gain M replaces A, R, Cpmax , and in

    (6). The air density is kept separate because air density is

    time varying and measurable.

    The control law (12) is defined separately for positive

    and negative regions of the rotor speed because it is

    undesirable to apply torque control when the turbine is

    spinning in reverse. Reverse operation can cause excessive

    wear on components that are designed for operation in onedirection.

    The equations for the gain adaptation law are

    M (k) = M (k 1) + M (k) , (13)

    M(k) = M sgn [M(k 1)] sgn[Pfavg(k)]

    |Pfavg(k)|1/2, (14)

    Pfavg(k) = Pfavg(k) Pfavg(k 1), (15)

    where k denotes the adaptive controllers discrete time

    step. The fractional average power Pfavg, given by

    Pfavg(k) =

    1n

    ni=1

    Pcap((k 1)n+ i )

    1n

    n

    i=

    1

    Pwy((k 1)n+ i )

    , (16)

    is the ratio of the mean power captured to the mean wind

    power. Pfavg is computed at each adaptive control time step

    k, where k is incremented once every n steps of region 2 oper-

    ation at the discrete-time torque control rate fs =

    100 Hz. Pwy, computed at 100 Hz, is the wind power given by

    Pwy =1

    2Av3 (cos )3 , (17)

    where is the yaw error, that is, the error between the

    wind direction and the yaw position of the turbine. Pcap is

    the captured power, given by

    Pcap = c +J, (18)

    which is also computed at 100 Hz. The yaw error factor

    (cos )3 in (17) shows that yaw errors reduce the power

    available to the turbine. The term c in the captured power

    Pcap is the generator power while J is the kinetic power

    (that is, the time derivative of the kinetic energy) of the rotor.

    In (13), M is adapted after n time steps of 10-ms periods

    of operation in region 2. Testing on CART indicates that

    the adaptation period must be on the order of hours; con-

    sequently, n = 1,080,000 steps, which corresponds to 3 h,

    is used in many CART experiments. This long time period

    is required in part because of the difficulty of obtaining ahigh correlation between measurements of wind speed

    over the entire swept area of the rotor and at the

    anemometer, which can be located either on the turbines

    nacelle or on a separate meteorological tower [16]. Another

    reason for the long adaptation period is that, since the tur-

    bine changes speed at a much slower rate than the wind,

    the slow responses must be averaged over time.

    In (14), the factor |Pfavg(k)|1/2 indicates the closeness of

    the adaptive gain M to its optimal value M, the gain that

    results in maximum energy capture. As M moves toward

    the peak of the curve in Figure 6, a given adaptation step

    M results in a smaller |Pfavg|because (dPfavg)/(d M) 0

    as M 0. Thus, |M| decreases as the optimal gain isapproached. The exponent 1/2 is chosen based on simula-

    tion, and selection of M > 0 is discussed below.

    In (16), Pcap is used rather than the rotor aerodynamic

    power P given by (3) because the sensor requirements for

    Pcap are more consistent with the instrumentation normal-

    ly available on industrial turbines. The two definitions of

    turbine power are closely related, differing only by the

    mechanical losses in the turbines gearbox; these losses

    make Pcap < P by a small amount.

    JUNE 2006 IEEE CONTROL SYSTEMS MAGAZINE 75

  • 8/3/2019 Control of Variable-speed Wind Turbines

    7/12

    Figure 6 portrays the output of constant-wind-speed

    simulations using the rigid body model (7) and the control

    torque (12). The model and controller are simulated with

    26 different values of the gain M, where each simulation

    lasts 200 s with constant M for the duration of the simula-

    tion. The turbines power output for each of the 26 gain

    values is averaged over each 200-s simulation to produce

    the solid Pfavg curve in Figure 6. In Figure 6, M = 174.5 is

    the optimal gain based on the standard torque control

    coefficient K in (6) as well as the simulated power-

    coefficient Cp surface in Figure 4. Since these data are

    obtained from simulations, the optimal gain M is known.

    The error M inM is given by

    M = M M.

    The adaptive controller attempts to have the turbine power

    track the wind power, assuming that the maximum power

    coefficient Cpmax and the optimal tip-speed ratio are

    unknown. In contrast, adaptive controllers such as those in

    [10][11] focus on different uncertainties and assume some

    knowledge of theCp surface, particularly andCpmax . In addi-

    tion, the averaging period used in this article is long compared

    to the time periods used by the adaptive controller in [9].

    Figure 7 shows data collected in the first year of adaptive

    CART operation. Only region 2 data is plotted, and the

    change in the adaptation period length from 10 min to 180

    min is apparent. The adaptation behavior with the longer

    adaptation period is significantly better than the behavior

    with the shorter adaptation period. The three discontinuities

    in the data reflect occasions where the adaptive controllerwas restarted due to a change in the method for calculating

    Pfavg and problems with sensors on CART. The last dozen

    adaptations oscillate about a value that is just less than 50%

    of the predicted optimal value M+ = 174.5 computed from

    the PROP model of CART. In comparison, the CART study

    [17], obtained with the turbine running in constant speed

    mode, gives a true optimal gain M around 47% of the pre-

    dicted optimal valueM+. The experimental results shown in

    Figure 7 indicate that modeling tools such as PROP [13] can

    lead to large errors in predicting the optimal value of the

    gain M. We now proceed with the stability analysis.

    STABILITYWe now consider the stability of the closed-loop system

    with the adaptive torque gain control law. Some of the

    results in this section appear in [18]. Although control of

    CARTs torque is a discrete-time problem, we simplify the

    stability analyses of the torque control law (12) by assum-

    ing that the torque control is continuous time. This simpli-

    fication is valid because the control time step of 0.01 s is

    much smaller than the tip-speed ratios time constant,

    which depends on wind speed [19] and is about 48 s for

    CART operating in region 2 wind speeds of 612 m/s.

    Also, we assume that the adaptive control gain M> 0 is

    constant in the torque control law (12) analysis; this

    assumption is valid because the gain adaptation takesplace discretely and on a time scale several orders of mag-

    nitude slower than changes in the wind speed and rotor

    speed (hours versus seconds). Thus, each result that is

    based on a constant M assumption holds for the duration

    of each 3-h adaptation period. Furthermore, M is con-

    strained to be positive since the control torque (12) cannot

    be negative. In all of these proofs, the air density is

    assumed to be a positive constant. In reality, changes in air

    density are small, typically not much greater than 5%. A

    76 IEEE CONTROL SYSTEMS MAGAZINE JUNE 2006

    FIGURE 7 Adaptive gain Mnormalized by the predicted optimal gain

    M+ during region 2 operation of CART. Discontinuities indicate

    restarts of the gain adaptation law due to changes in the law and

    turbine sensor errors. In the second half of the data, Moscillates

    around the value 0.47 M+, which is approximately equal to the true

    optimal torque gain M.

    1.5

    1

    0.5

    NormalizedM(M/M+)

    00 20 40

    Time (h)

    60 80

    FIGURE 6 Pfavg versus M for the CART model. Pfavg is the ratio of

    the mean captured power to the mean wind power, while M is the

    error between the torque control gain Mand its optimal value M.

    The shape of this curve is based on the shape of CARTs Cp

    curve. In the adaptive controller, the gain adaptation law converges

    in part due to the shape of the Pfavg curve.

    100 50 0 50 1000.30

    0.32

    0.34

    0.36

    0.38

    0.40

    0.42

    Pfavg

    Gain Error M= M* M

    FractionalAve

    ragePowerPfavg

  • 8/3/2019 Control of Variable-speed Wind Turbines

    8/12

    simplified block diagram for these continuous-time sys-

    tems is given in Figure 8(a), where the linear plant is given

    by (7) and the nonlinear controller is given by (12).

    Asymptotic Stability of Zero Rotor SpeedFirst, we consider the asymptotic stability of the rotor-

    speed equilibrium = 0 in the absence of wind and in

    constant wind. To minimize energy loss in wind turbines,

    friction and drag due to mechanical bearings, gear mesh,

    generator core losses, and air resistance are designed to be

    as small as possible. However, in the analysis of asymptot-

    ic stability of the equilibrium point = 0, we revise (7) so

    that the angular acceleration includes a damping term

    b, where the damping coefficient b > 0, which yields

    = 1J(aero c b). (19)

    Using (8) and (12), (19) can be expanded to

    =

    12JARCqv2 bJ , < 0,

    12JARCqv

    2 J M

    2 bJ , 0.(20)

    Theorem 1

    Suppose that the wind speed v = 0 and M > 0 are con-

    stant. Then the equilibrium = 0 of the closed-loop sys-

    tem (20) is asymptotically stable.

    Proof

    For the initial condition (0) = 0 , the solution to (20)

    when v = 0 is

    ( t) =

    0e b

    J t, < 0,0b

    (b+M0)ebJtM0

    , 0.

    Hence, 0 as t .

    We also note that when the damping coefficient b = 0

    and the wind speed v = 0, (20) becomes

    =

    0, < 0,

    JM

    2, 0,

    which has the solution

    ( t) =

    0, < 0,JMt+ J

    0

    , 0.

    In this case, 0 holds only when the rotor is spinning in the

    positive direction, which is normal operation for the turbine.

    Asymptotic Stability of Rotor Speed

    with Constant, Positive Wind InputThe next stability result concerns the convergence of the

    rotor speed to an equilibrium value under an idealized

    constant, positive wind speed. This analysis is similar to

    the one describing Figure 5 and given in (5)(11). Once

    again, the plant is given by (19) and the nonlinear con-

    troller is given by (12). The adaptive controller (12) does

    not assume knowledge of the aerodynamic parameters

    Cpmax and . Setting the 0 portion of (20) equal to zero

    and solving for Cp in terms of using (4) and (9) yields

    Cp =M3v + 2bR

    12 AR

    3v G(,M, b, v). (21)

    The equilibrium points = 0 of turbine operation are thus

    given by the intersection of with the turbines Cp

    curve. Figure 9 shows CARTsCp curve and two illus-

    trative G(,M, b, v) curves plotted using representative

    values of , v, and b.

    In Figure 9, the cubic functions G(,M, b, v) do not inter-

    sect the Cp curve at the peak of the curve when the adaptive

    JUNE 2006 IEEE CONTROL SYSTEMS MAGAZINE 77

    FIGURE 8 Control loops for (a) the aerodynamic torque aero and

    rotor speed and (b) the gain adaptation law. (a) Stability of the

    continuous-time control loop is analyzed by Theorems 13, while

    Theorem 4 considers (b) the discrete-time adaptive loop.

    +

    NonlinearController

    LinearPlant

    c

    aero

    +

    NonlinearController

    PfavgNonlinearPlant

    M

    M*

    (a) (b)

    FIGURE 9 CARTs power coefficient Cp curve and cubic functions

    for two values of the adaptive gain M. When M is not equal to its

    optimum value M, the intersection of the Cp and G(, M) curves

    does not occur at the peak of the Cp curve, which leads to subopti-

    mal energy capture. Similar to Figure 5, the intersection of each

    cubic curve with the Cp curve is an equilibrium point of the system

    for the indicated adaptive gain M.

    2 4

    0.4

    0.3

    0.5

    0.2

    0.1

    06 8

    Tip-Speed Ratio

    PowerCoefficientCp

    10 12 14

    G( ,M)M= 1.3M*

    G( ,M)M= 0.7M*

    Cart CpVersus

  • 8/3/2019 Control of Variable-speed Wind Turbines

    9/12

    gain M = M ; thus, the equilibrium point of the system is

    suboptimal in terms of energy capture. Let 2 be the highest

    value of for which the curve G(,M) intersects Cp().

    Mathematically, 2 is the tip-speed ratio for which

    G(,M) > Cp() for all > 2. Let 1 denote the next highest

    intersection point, that is, the value of for which

    0 < 1 < 2 and G(,M) < Cp() for all 1 < < 2 and

    G(,M) > Cp() for all < 1 within a neighborhood of 1.

    For the dashed curve M = 0.7M in Figure 9, these values

    correspond to 1 = 3.1 and 2 = 8.4. The following result

    shows that, for a constant wind input, the tip-speed ratio con-

    verges to 2 as long as the initial value of is greater than 1.

    Theorem 2

    Suppose that the wind speed v and the adaptive gain M are

    positive constants and 1 > 0. Then the equilibrium point

    = 2 of the closed-loop system consisting of the plant

    =R

    Jvaero c

    bv

    R (22)

    and the nonlinear controller (12) is locally asymptotically

    stable with domain of attraction (1,).

    Proof

    First note that > 0 for all 0 < 1 < since = v/R from

    (4). Define = 2 and the Lyapunov candidate

    V= (1/2)2 . For > 0,

    V= ( 2)

    1

    2JAR2Cqv

    1

    JRM2v

    b

    J

    = ( 2)h(Cq, v, ).

    (23)

    Substituting Cp/ for Cq in (23) and applying (21) yields

    h(Cq, v,) > 0 for all such that 1 < < 2 , that is,

    G(,M) < Cp(). Moreover, > 2 givesG(,M) > Cp()by

    definition of 2, and therefore h(Cq, v,) < 0by definition (21)

    of G. Thus, V< 0 for all (1,) except = 2, for which

    V= 0. Hence, the equilibrium point = 2 of (22) is locally

    asymptotically stable. Finally, it is easy to show that the

    domain of attraction is (1,). Note that V is bounded away

    from zero on every connected, compact subinterval of (1,)

    that does not contain 2. Thus, the time required for to reach

    the edge of the subinterval closest to 2 is finite. Now, moves

    monotonically toward 2. If does not converge to 2, then thetime it takes to reach the edge closest to 2 of a subinterval

    not containing 2 must be infinite, which contradicts the earlier

    result. Thus, the domain of attraction is (1,).

    The convergence of the tip-speed ratio to 2 is equiva-

    lent to the convergence of the rotor speed to 2v/R for a

    specific wind speed v. Furthermore, when M = M , the

    curves G(,M) and Cp() intersect at (,Cpmax ) as shown

    for the standard torque control in Figure 5; therefore, optimal

    energy capture is achieved for the constant wind input case.

    We acknowledge that zero and constant wind speeds

    never occur in the field. However, wind speeds near zero do

    occur during turbine operation, causing a shutdown when

    the wind speed is close to zero for a sufficiently long time.

    These results are useful for developing an understanding of

    the torque control law, although the cases are idealized.

    Input-Output StabilityNext, we show that a bounded input (squared wind speed

    v2) to the system produces a bounded output (rotor speed

    ). All wind turbines have a maximum safe operating

    speed, and often pitch control is used to prevent the tur-

    bine from operating at speeds above this maximum. Nev-

    ertheless, an enhanced understanding of the wind turbine

    control system can be achieved by examining whether the

    torque control (12) bounds the turbine speed. The follow-

    ing result considers a time-varying wind speed v.

    For T> 0, we use the standard the definition of the L2norm of v(t) given by

    vL2[0,T] =

    T0v(t)2dt.

    Theorem 3

    Suppose the rotor torque coefficient Cq 1, the adaptive

    gain M > 0 is constant, and consider the closed-loop tur-

    bine system (12), (19) with input given by the squared wind

    speed v2 and output given by the rotor speed . Then, for

    all finite T> 0, the system (12), (19) is L2 stable on [0,T].

    Proof

    Consider the kinetic energy EK = (1/2)J2 of the rotor

    and define

    V=1

    ARJ2, (24)

    where > 0 is a constant. The time derivative of (24) is

    V=

    Cqv2

    b12 AR

    2, < 0,

    Cqv2 b1

    2AR

    2 M12AR

    3, 0.(25)

    Le t = b/((1/2)AR) . Then, since Cq 1 an d

    (M/((1/2)AR))3 0 for 0, it follows that

    V v2 2. (26)

    Thus, Lemma 6.5 in [20] implies that the wind turbine sys-

    tem, from the squared wind speed v2 to the rotor speed ,

    is finite-gain L2 stable over [0,T].

    The restriction that T be finite is necessary due to the

    nature of the wind speed v(t). Since wind speed v(t) > 0

    can hold at all times, it is possible that v(t) / L2[0,].

    Thus, Tmust be finite to guarantee that proof of L2 stability

    in [0,T] makes sense.

    78 IEEE CONTROL SYSTEMS MAGAZINE JUNE 2006

  • 8/3/2019 Control of Variable-speed Wind Turbines

    10/12

    The condition Cq 1 is usually satisfied for modern

    turbines in normal region 2 operation. The Betz limit [14],

    which is the theoretical maximum power coefficient Cp for

    any real turbine, has a value of Cp = 16/27. Since

    Cq = Cp/ [see (9)], it follows that Cq 1 for 16/27.

    When 16/27, it follows from the definition of tip-speed

    ratio in (4) that = v/R (16/27)v/R.

    For finite > 0 and [0,T],L , that is, bounded

    input, bounded output stability of with respect to the

    input v is given by = v/R.

    Theorem 3 shows that a wind turbine is not a perpetu-

    al motion machine. Since the assumption that M is con-

    stant holds only for the duration of an adaptation period,

    Theorem 3 shows that the energy produced by a turbine

    is less than that contained in the wind over each adapta-

    tion period.

    Convergence of the Gain Adaptation AlgorithmThe final stability analysis examines convergence of the

    adaptive gain M

    M

    using the gain adaptation law(13)(15). Figure 8(b) shows a simplified block diagram

    for this system, where the nonlinear plant is the fractional

    average power Pfavg versus torque gain error M relation-

    ship shown in Figure 6 and the nonlinear controller is

    given by (13)(15).

    We make two assumptions before studying the stability

    properties of the gain adaptation law.

    Assumption 1

    The optimum torque control gain M is constant.

    The turbines aerodynamic parameters, and thus M ,

    change with time due to blade erosion, residue buildup,

    and related events. However, we can assume that M

    isconstant because the turbines physical changes are typical-

    ly noticeable only over months or years, whereas the gain

    adaptation law has an adaptation period of less than a day.

    Assumption 2

    The Pfavg versus M curve has a maximum at M = 0, is con-

    tinuously differentiable, and is strictly monotonically

    increasing on M < 0 and strictly monotonically decreasing

    on M > 0. Experimental data [17] support this assumption.

    For the initial conditions M0 , Pfavg0, M0 , and M1 ,

    k > 2 is the time frame of interest in the convergence analy-

    sis. Theorem 4 covers only the time k > 2because the first

    two steps are more influenced by the initial guesses than bythe turbines aerodynamic properties.

    We begin the convergence analysis by considering how

    the adaptive gain can diverge, that is, | M| as k .

    One possibility is | Mk| > | Mk1| with either sgn( Mk) = 1

    or sgn( Mk) = 1 for all k > 2. However, it is easy to show

    that this scenario cannot occur with the gain adaptation

    law (13)(15). Indeed, the adaptive torque gain error M

    cannot take two consecutive steps in the wrong (incorrect)

    direction for all k > 2, as shown by the following result.

    Theorem 4

    Let k > 2. Under Assumptions 1 and 2 and the gain adap-

    tation law (13)(15), | Mk+1| > | Mk| > | Mk1| never occurs

    when sgn( Mk+1) = sgn( Mk) = sgn( Mk1).

    Proof

    Suppose Mk+1 > Mk > Mk1 and sgn( Mk+1) = sgn( Mk) =

    sgn( Mk1) = 1 for some k > 2. Note that Mk > Mk1 gives

    Mk Mk1 = Mk > 0, (27)

    which implies that Mk < 0. Furthermore, Mk+1 > Mk gives

    Mk+1 Mk = Mk+1 > 0, (28)

    which implies that Mk+1 < 0. By (16)(18), Pfavg k+1 is cal-

    culated at the end of the adaptation interval during which

    M = Mk ; thus, Pfavg k+1 is calculated from data collected

    while the torque gain error was Mk . Since Mk > Mk1 ,

    Assumption 2 implies Pfavg k+1 < Pfavg k . Therefore, by (15),

    Pfavg k+1 < 0. (29)

    In (27) and (29), sgn(Mk) = sgn(Pfavg k+1 ) = 1. Thus, by

    (14), sgn(Mk+1) = 1, contradicting (28). Thus, it is

    impossible for both Mk+1 > Mk > Mk1 and sgn

    ( Mk+1) = sgn( Mk) = sgn( Mk1) = 1 to be true. A similar

    argument can be used for negative values of M.

    Since the sign of the adaptation step M cannot be

    incorrect for two consecutive steps, the gain M, which

    affects the magnitude of M, is the critical factor in deter-

    mining whether the adaptive gain diverges. Figure 10

    shows an example in which the gain M is large enoughto cause the adaptive gain M to diverge. In this example,

    | Mk+1| > | Mk1| for all k > 2, although both | Mk+1| > | Mk|

    and | Mk+1| < | Mk| occur when k > 2.

    JUNE 2006 IEEE CONTROL SYSTEMS MAGAZINE 79

    FIGURE 10 Adaptive gain steps in an unstable case. The numbers

    19 indicate the discrete-time steps. In this case, the gain M in

    the gain adaptation algorithm (13)(15) is too large, and thus the

    gain adaptation law diverges.

    20 1015 5 0 105

    9

    5

    1

    7

    36 2 48

    20150.25

    0.30

    0.35

    0.40

    0.45

    0.50

    Gain Error M= M* M

    FractionalA

    veragePowerPfavg

  • 8/3/2019 Control of Variable-speed Wind Turbines

    11/12

    Since M diverges if |Mk| > | Mk1| for all k > 2, we

    consider

    |Mk| = | Mk1|, Mk1 = 0 (30)

    to be the critical case, or the marginal stability case. Defineykby

    yk a M2k1 + P0, (31)

    whereyk is a curve satisfying Assumptions 12 whose form

    is better known than Pfavgk . In (31), a < 0 and P0 is a real

    number; (30) can be solved for the critical gain M. For

    consistency with the discrete-time indices in the equation(16) for Pfavgk , yk is a function of

    Mk1 rather than of Mk.

    In the critical gain scenario of this example, the system

    alternates among the three points plotted in Figure 11. If

    Mk = Mk1 , then the error Mk = 0 by (13). Substituting ykfor Pfavgk in (15) and considering (14), the gain M is such

    that Mk+1 = Mk, resulting in Mk+1 = Mk1. Following

    the equations through one more step shows that Mk+2 = 0,

    and the adaptive gain alternates among these three points.

    Thus, an upper bound on the gain M for stability can be

    found by equating

    Mk = Mk1 = Mk+1

    and solving for M in terms of a with Mk = 0, which yields

    M =

    1|a| . (32)

    Thus, if 0 < M < |a|1/2 , then the gain adaptation law

    (13)(15) does not cause divergence of the adaptive

    torque control gain error M on the curve (31). In fact,

    since M =|a|

    1/2

    is the marginal stabil i ty case,0 < M < |a|1/2 yields M 0. Since this bound on M

    depends on the magnitude |a|, every gain M chosen

    for a given value of a in (31) also guarantees conver-

    gence of the adaptive gain M on a curve with a smaller

    value of a.

    We can state a similar result for a curve that is not even

    [as in (31)], that is, one for which Pfavg( M) = Pfavg( M)

    does not hold. If the gain M is chosen to guarantee con-

    vergence based on the slope of the steeper side of the

    curve, then M guarantees convergence over the entire

    curve. Thus, for an arbitrary Pfavg versus M curve, there

    exists M > 0 that guarantees convergence of the adap-

    tive gainM, and this gain M depends on the steepness ofthe Pfavg versus M curve.

    Since there are no turbines for which the Pfavg versus M

    curve is well known, an approximation of the curve is nec-

    essary to control each turbine. The more conservative the

    choice of M, the more likely it is that M converges to M

    since the gain adaptation law (13)(15) is more robust to

    errors in the approximated Pfavg versus M curve for small-

    er M. However, a smaller M also results in smaller

    step sizes and thus might cause the convergence to occur

    more slowly.

    An example of the choice of M is provided in Figure

    12. The coefficients a and P0 of (31) are chosen so that (31)

    fits snugly inside the Pfavg curve, being coincident atM= 0 and satisfying y < Pfavg for M = 0. In this case,

    a= 0.00001 m10 . Thus, the maximum allowable gain

    M for stability is 316 m5. The gain used in testing on

    CART before this stability analysis was performed was

    M = 100 m5, which was determined empirically from

    simulations and early hardware testing. Although actual

    turbine results indicate stable performance of the adaptive

    control law, this stability analysis provides further reassur-

    ance and guidelines in choosing M.

    80 IEEE CONTROL SYSTEMS MAGAZINE JUNE 2006

    FIGURE 12 CART Pfavg versus Mcurve and symmetric inset curve.

    The curve labeled Pfavg is identical to the curve shown in Figure 6,

    while the quadratic curve labeled yis added to illustrate the method

    for selecting the adaptive gain M. When the quadratic curve is

    chosen such that y(M) Pfavg(M) and y(M) = Pfavg(M) if and only

    if M= 0, the upper limit on M for stability of the gain adaptation

    law is a function of the coefficient of the squared term in (31).

    100 50 0 50 1000.30

    0.32

    0.34

    0.36

    0.38

    0.40

    0.42Pfavg

    y

    Gain Error M= M* M

    Fra

    ctionalAveragePowerPfavg

    FIGURE 11 Finding the critical gain M. Marginal stability of the

    gain adaptation law, defined as oscillation among three points on

    the ycurve (31), occurs when the step size Mk has the same

    magnitude as the error Mk1 for a symmetric curve.

    50 0 50

    0.40

    0.41

    0.42

    0.43

    Gain Error M= M* M

    (Mk,yk+1)

    MkMk+1

    FractionalAveragePowerPfavg

    (Mk+1,yk+2)

    (Mk1,yk)

  • 8/3/2019 Control of Variable-speed Wind Turbines

    12/12

    CONCLUSIONS

    This article considers an adaptive control scheme previously

    developed for region 2 control of a variable-speed wind tur-

    bine. In this article, we addressed the question of theoretical

    stability of the torque controller, showing that the rotor speed

    is asymptotically stable under the torque control law (12) in the

    constant wind speed input case and L2 stable with respect to

    time-varying wind input. Further, we derived a method for

    selectingM in the gain adaptation law (13)(15) to guarantee

    convergence of the adaptive gainM to its optimal valueM.

    ACKNOWLEDGMENTS

    This work was supported in part by the U.S. Department of

    Energy through the National Renewable Energy Laboratory

    under contract DE-AC36-99G010337, the University of Col-

    orado at Boulder, and the American Society for Engineering

    Education. We would also like to acknowledge Prof. Dale

    Lawrence and Dr. Vishwesh Kulkarni for their suggestions

    on improving our article.

    AUTHOR INFORMATION

    Kathryn E. Johnson ([email protected]) received the B.S.

    degree in electrical engineering from Clarkson University in

    2000 and the M.S. and Ph.D. degrees in electrical engineering

    from the University of Colorado in 2002 and 2004, respective-

    ly. In 2005, she completed a postdoctoral research assignment

    studying adaptive control of variable-speed wind turbines at

    the National Renewable Energy Laboratorys National Wind

    Technology Center. That fall, she was appointed Clare Boothe

    Luce Assistant Professor at the Colorado School of Mines in

    the Division of Engineering. Her research interests are in con-

    trol systems and control applications. She can be contacted at

    Colorado School of Mines, Division of Engineering, 1610 Illi-nois St., Golden, CO 80401 USA.

    Lucy Y. Pao received the B.S., M.S., and Ph.D. degrees

    in electrical engineering from Stanford University. She is

    currently a professor of electrical and computer engineer-

    ing at the University of Colorado at Boulder. She has pub-

    lished over 120 journal and conference papers in the area

    of control systems. Her awards include the Best Commer-

    cial Potential Award at the 2004 International Symposium

    on Haptic Interfaces for Virtual Environments and Teleop-

    erator Systems as well as the Best Paper Award at the 2005

    World Haptics Conference. She was the program chair for

    the 2004 American Control Conference, and she is current-

    ly an elected member on the IEEE Control Systems SocietyBoard of Governors.

    Mark J. Balas has made theoretical contributions in

    linear and nonlinear systems, especially in the control of

    distributed and large-scale systems, aerospace structure

    control, and variable-speed, horizontal-axis wind turbine

    control for electric power generation. He is a Fellow of

    the IEEE and the AIAA. He is currently head of the Elec-

    trical and Computer Engineering Department at the Uni-

    versity of Wyoming.

    Lee J. Fingersh received the B.S. and M.S. degrees in

    electrical engineering from the University of Colorado in

    1993 and 1995, respectively. He has been employed at

    NREL since 1993, working in the fields of aerodynamics

    testing, power electronics, electric machines, energy stor-

    age, and controls. Most recently, he has been responsible

    for a large controls field testing project and its associated

    test machine, the Controls Advanced Research Turbine.

    REFERENCES[1] Global wind energy installations climb steadily, American Wind Ener-

    gy Associations Wind Power Outlook 2005 [Online], Mar. 2004, p. 6. Avail-able: http://www.awea.org/pubs/documents/Outlook%202005.pdf[2] K. Johnson, L. Fingersh, M. Balas, and L. Pao, Methods for increasing

    region 2 power capture on a variable speed wind turbine, J. Solar Energy

    Eng., vol. 126, no. 4, pp. 10921100, 2004.[3] J. Svensson and E. Ulen, The control system of WTS-3 instrumentationand testing, in Proc. 4th Int. Symp. Wind Energy Systems , Stockholm,

    Sweden, 1982, pp. 195215.[4] L. Fingersh and P. Carlin, Results from the NREL variable-speed testbed, in Proc. 17th ASME Wind Energy Symp.}, Reno, NV, 1998, pp. 233237.

    [5] K. Stol and M. Balas, Periodic disturbance accommodating control forspeed regulation of wind turbines, in Proc. 21st ASME Wind Energy Symp.,

    Reno, NV, 2002, pp. 310320.[6] A.Eggers, H. Ashley, K. Chaney, S. Rock, and R.Digumarthi, Effects of cou-pled rotor-tower motions on aerodynamic control of fluctuating loads on light-weight HAWTs, in Proc. 17th ASME Wind Energy Symp., Reno, NV, 1998, pp.

    113122.[7] M. Hand, Load mitigation control design for a wind turbine operatingin the path of vortices, in Proc. Science of Making Torque from Wind 2004 Spe-cial Topic Conf., Delft, The Netherlands, 2004 [CD-ROM].[8] A. Wright and M. Balas, Design of controls to attenuate loads in theControls Advanced Research Turbine, J. Solar Energy Eng., vol. 126, no. 4,

    pp. 10831091, 2004.[9] S. Bhowmik, R. Spe, and J. Enslin, Performance optimization for dou- bly-fed wind power generation systems, IEEE Trans. Ind. Applicat., vol. 35,

    no. 4, pp. 949958, 1999.[10] J. Freeman and M. Balas, An investigation of variable speed horizontal-axis wind turbines using direct model-reference adaptive control, in Proc.18th ASME Wind Energy Symp., Reno, NV, 1999, pp. 6676.

    [11] Y. Song, B. Dhinakaran, and X. Bao, Variable speed control of wind tur-bines using nonlinear and adaptive algorithms, J. Wind Eng. Ind. Aerodyn.,

    vol. 85, no. 3, pp. 293308, 2000.[12] M. Simoes, B. Bose, and R. Spiegel, Fuzzy logic based intelligent con-trol of a variable speed cage machine wind generation system, IEEE Trans.Power Electron., vol. 12, no. 1, pp. 8795, 1997.[13] S. Walker and R. Wilson, Performance Analysis Program for Propeller TypeWind Turbines. Corvallis, OR: Oregon State Univ., 1976.[14] T. Burton, D. Sharpe, N. Jenkins, and E. Bossanyi, Wind Energy

    Handbook. New York: Wiley, 2001.[15] D. Simms, S. Schreck, M. Hand, and L. Fingersh, NREL unsteady aero-dynamics experiment in the NASA-Ames wind tunnel: A comparison of

    predictions to measurements, NREL Rep. TP-500-29494, 2001 [Online].Available: http://www.nrel.gov/docs/fy01osti/29494.pdf.[16] K. Johnson, L. Fingersh, L. Pao, and M. Balas, Adaptive torque control

    of variable speed wind turbines for increased region 2 energy capture, inProc. 2005 ASME Wind Energy Symp., Reno, NV, 2005, pp. 6676.

    [17] L. Fingersh and K. Johnson, Baseline results and future plans for theNREL controls advanced research turbine, in Proc. 23rd ASME Wind Energy

    Symp., Reno, NV, 2004, pp. 8793.[18] K. Johnson, L. Pao, M. Balas, V. Kulkarni, and L. Fingersh, Stability

    analysis of an adaptive torque controller for variable speed wind turbines,in Proc. IEEE Conf. on Decision and Control, Atlantis, Paradise Island,Bahamas, 2004, vol. 4, pp. 40164021.

    [19] K. Pierce, Control method for improved energy capture below ratedpower, Proc. ASME/JSME Joint Fluids Engineering Conf., San Francisco, CA,1999, pp. 10411048.

    [20] H. Khalil, Nonlinear Systems, 3rd ed. Upper Saddle River, NJ: Prentice-Hall, 2002, p. 242.[21] P. Gipe, Wind Energy Comes of Age. New York: Wiley, 1995.

    JUNE 2006 IEEE CONTROL SYSTEMS MAGAZINE 81