You are on page 1of 19

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/268470595

State-Space Control of Tower Motion for Deepwater Floating Offshore Wind Turbines

Conference Paper · January 2008


DOI: 10.2514/6.2008-1307

CITATIONS READS

19 369

3 authors, including:

Karl Alexander Stol


University of Auckland
95 PUBLICATIONS   1,210 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Wind Disturbance Rejection View project

All content following this page was uploaded by Karl Alexander Stol on 17 June 2015.

The user has requested enhancement of the downloaded file.


46th AIAA Aerospace Sciences Meeting and Exhibit AIAA 2008-1307
7 - 10 January 2008, Reno, Nevada

State-Space Control of Tower Motion for Deepwater


Floating Offshore Wind Turbines

Hazim Namik * and Karl Stol †


Department of Mechanical Engineering, The University of Auckland, Auckland, New Zealand

and

Jason Jonkman ‡
National Renewable Energy Laboratory, Golden, Colorado, 80401-3393

The offshore wind energy potential is huge and to capture that energy, turbines have to
be placed further offshore. Floating wind turbines offer a solution for deep waters. However,
a floating wind turbine has extra motions that will affect the turbine in power production
and structural loads. Therefore, the turbine control system has to be able to regulate power
production and maintain safe operation of the turbine under incident wind and wave
conditions. The work presented here discusses the application of state-space optimal
controllers to regulate rotor speed and platform pitch above rated wind speed. A gain
scheduled PI controller is used as a baseline to gauge the performance of the new controllers
developed. A collective pitch linear quadratic regulator, designed to only regulate rotor
speed and platform pitch, improves system performance but this improvement is thought to
be due to better controller tuning as both controllers use the same mechanism to restore
platform pitch and regulate speed. Individual blade pitch control using periodic control
theory is applied because it uses a different mechanism to regulate platform pitch.
Preliminary simulation results show that individual blade pitch control has platform pitch
regulation over collective pitch controllers. However, unintended excitation of platform roll
indicate that a more complicated controller may be required to ensure closed loop stability
of the entire floating turbine.

Nomenclature
A = state matrix
AAvg = azimuth averaged state matrix
A(t) = periodic state matrix
B = actuator gain matrix
BAvgB = azimuth averaged actuator gain matrix
B(t) = periodic actuator gain matrix
C = measurement matrix corresponding to the states
CP,Max = maximum power coefficient
D = measurement matrix corresponding to actuator inputs
G(t) = periodic gain matrix
J = solution of the quadratic cost function
K = region 2 generator torque controller constant of proportionality
KP, KI = proportional and integral gains
KLQR = linear quadratic regulator gain matrix
N = gearbox ratio
*
PhD Candidate, Department of Mechanical Engineering, 20 Symonds Street, Auckland, Private Bag 92019,
Auckland Mail Centre, Auckland 1142, New Zealand, AIAA Member.

Senior Lecturer, Department of Mechanical Engineering, 20 Symonds Street, Auckland, Private Bag 92019,
Auckland Mail Centre, Auckland 1142, New Zealand, AIAA Member.

Senior Engineer I, National Wind Technology Center, 1617 Cole Boulevard, AIAA Member.

1
American Institute of Aeronautics and Astronautics

Copyright © 2008 by H. Namik, K. Stol, and J. Jonkman. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
P = solution of the algebraic Riccati equation
P(t) = solution of the periodic Riccati equation
PRated = rated power of the wind turbine
Q = states weighting matrix
R = actuators weighting matrix
RRotor = rotor radius
TGen = applied generator torque
t = time
u = control inputs vector
x = states vector
ηGen = generator efficiency
θ = blade pitch
λ = tip speed ratio
ρ = air density
φ = platform pitch angle
ψ = rotor azimuth angle
ωGen = generator speed

I. Introduction

W IND energy is a clean, renewable, and sustainable source of energy. Today, power generated from wind
energy is the fastest growing energy source in the world1. A recent trend in the wind turbine industry is to go
offshore, mainly due to better wind resource attributes when compared to onshore wind. Offshore wind is stronger
and steadier with less turbulence than onshore wind2-4. Offshore wind also has less vertical shear meaning higher
wind speed at lower hub heights and is more spatially consistent than onshore. On average, the annual mean
offshore wind speed is 20% higher than onshore2.
The offshore wind energy potential is huge. The UK offshore potential is estimated to be 986TWh/year while
demand is estimated to be 321TWh/year; i.e. more than three times the demand2. The US offshore potential for
winds between 10km and 100km offshore is estimated to be more than 900GW which is more than currently
installed generation capacity; however, much of this potential is located in deep water, deeper than 30m5.
Currently, the deepest installed offshore wind turbine is the Beatrice wind farm off the shore of Scotland6; it is
constructed in 44m deep water. However, for water deeper than 60m the only feasible option is to have floating
wind turbines4, 7; these floating turbines are sometimes referred to as deepwater floating wind turbines. One major
disadvantage of floating wind turbines is that motions induced by wind and wave conditions are almost impossible
to eliminate thus the design of these turbines must take into account the added degrees of freedom due to platform
motions.
There are three main floating concepts: a Spar-buoy, a Tension Leg Platform (TLP), and a Barge platform8. Each
uses a different floating principle as shown in Fig. 1. The work presented here will be based on the floating barge
concept for two reasons. First, the barge with catenary mooring lines is the least expensive option out of many
floating platform concepts9, 10. It is assumed that a good control system would minimize the effects of platform
motions on static loads such that current offshore turbines are used on any platform without changes to its design.
Therefore, the capital cost of the turbine is the same for all platform types thus the barge concept is the most feasible
given that such controllers can be designed. Second, the barge concept is well documented in the literature8, 9, 11-13. In
Ref. 5 and 8 a fully coupled aero-hydro-servo-elastic simulator is developed by Jonkman et al. for a generic floating
offshore wind turbine. In Ref. 9 and 11 a model for the floating barge concepts is developed, Proportional-Integral
(PI) controllers are tested on the barge concepts in Ref. 12, and loads analysis for extreme wind and wave conditions
are studied in Ref. 13.
Most of the literature about floating wind turbines is focused on the design of the platform and comparing
between different concepts; very little work has been done to design a control system for the floating turbines.
Neilsen et al.4, working on a spar-buoy floating concept, developed an active control strategy to avoid structural
resonance and take into account the fact that as the wind speed increases the thrust force decreases which may cause
negative damping of platform motions in above rated wind speeds. This negative damping can occur if the turbine is
pitching forward increasing the relative wind speed which will cause a standard controller to increase the blade pitch
to reduce aerodynamic torque and hence reducing the restoring thrust force. The controller introduced by Nielsen et
al. achieves satisfactory results in reducing the platform resonant motions in region 3 – above rated wind speed – in
simulation and scale model testing.

2
American Institute of Aeronautics and Astronautics
Figure 1. The three main floating concepts8

This study will continue the work done by Jonkman12 to design and test floating wind turbine controllers.
Jonkman has developed a gain scheduled PI blade pitch control coupled with a baseline torque controller to regulate
the turbine speed on a floating barge concept; this is referred to as the baseline controller in that study. The baseline
controller was then simulated using the newly developed fully coupled simulator. System performance had large
platform pitch oscillations resulting in large power fluctuations. Therefore, three modifications to the baseline
controller were developed; these include:
ƒ Feedback of tower top acceleration: Additional proportional control loop that aims to reduce tower fore-
aft motion based on the measurement of the tower top acceleration.
ƒ Active pitch to stall speed control: The aim here is to get the extra restoring thrust force once the blades
are stalled when the turbine pitches forward.
ƒ Detuning the controller gains: Detuning the gains will reduce the use of the blades and possibly reduce
the negative damping effect.
The addition of tower top feedback control loop does not improve platform damping; this is because of
conflicting controller demands. As the platform is pitching forward (into the wind), the PI speed controller pitches
the blades to feather (positive) to reduce torque while the tower feedback loop pitches the blades to fine pitch
(negative) to increase thrust. The active pitch to stall controller has excellent power regulation, however, platform
motions are not reduced but are increased instead. This contradictory result can be explained by looking at the open
and ideal closed loop damping ratios (not shown here). The study concludes that the pitch to stall damping is
actually positive. And, since the baseline pitch to feather controller had better damping, it can be concluded that the
actual closed loop damping of the baseline controller is greater than that of the pitch to stall12; i.e. the system has
positive damping. The detuned gains produce the best results out of the four controllers. Reducing the controller
gains makes the response of the system closer to that of the open loop response hence increase damping. This
produces reasonable power regulation and slightly reduces the platform oscillations.

3
American Institute of Aeronautics and Astronautics
The main objective of this study is to investigate whether using modern control theory such as using state-space
control theory can further reduce platform motions induced by incident wind and waves; minimizing platform
motions will help reduce bending loads on the turbine and the impact on power production. The premise is that
state-space controllers can handle multiple objectives and are Multi-Input-Multi-Output (MIMO) rather than Single-
Input-Single-Output (SISO), such as the controller implemented in Ref. 12. Therefore, state-space controllers can
intrinsically accommodate individual blade pitch control where each blade is commanded differently. Optimal
control theory can be used to optimize the controller gains to meet several objectives; a Linear Quadratic Regulator
(LQR) is an example of an optimal controller designed to minimize the cost function in Eq. (1) where J is the cost
function, x and u are the state and control input vectors respectively, Q and R are the weighting matrices associated
with the state and actuator vectors respectively. The objectives of the state-space controller are to regulate the
turbine rotor speed and platform pitch while ensuring stability of the turbine in above rated wind speed region
(region 3).

( )

J = ∫ x Q x + u Ru dt
T T
(1)
0

State-space controllers have been designed and implemented on onshore wind turbines with multiple objectives
successfully; two examples are given here. Wright14 has developed several state-space controllers with multiple
objectives to control a 600kW wind turbine. His most complicated controller involved a 9-state individual blade
pitch Disturbance Accommodating Controller (DAC) designed to regulate rotor speed, maintain stable closed loop
behavior, and reduce loads due to spatial variations induced by wind shear on the rotor. The controller was able to
achieve all of its objectives. Stol et al.15 were able to develop collective and individual blade pitch controllers to
reduce turbine loads as well as regulate the rotor speed using state-space techniques. They showed that a multi-
objective controller was able to reduce the loads when compared to a standard PI controller. Furthermore, they
showed that individual blade pitch controller was able to further reduce the loads than a collective pitch state-space
controller; however, field tests did not reflect that significant difference in simulation between collective and
individual blade pitch control.
In this paper, analysis tools and model description are discussed in Section II. Section III describes the collective
pitch state-space controller developed for the floating wind turbine on the barge while Section IV describes the
preliminary work done to implement an individual blade pitch controller. Finally, conclusions are presented in
Section V.

II. Floating Offshore Wind Turbine Analysis Tools

A. Floating Turbine Simulator


The simulation tools that will be used to design and analyze the developed controllers for the floating wind
turbine are FAST and MATLAB/Simulink. FAST (Fatigue, Aerodynamics, Structures and Turbulence) is a freely
available design code developed by the National Renewable Energy Laboratory (NREL)16; it is of moderate
complexity used to analyze the structural dynamics of horizontal axis wind turbines. FAST models the tower, blades
and the drive train as flexible elements and using assumed bending modeshapes 5, 16. Each blade has two flapwise
and one edgewise bending modes. The tower has two fore-aft and two side-side bending modes. The drive train
flexibility is modeled through a linear spring and a damper for the low speed shaft. The remaining elements of the
wind turbine (nacelle and hub) are modeled as rigid bodies. The fidelity of the model can be set by selecting which
degrees of freedom (DOF) are to be taken into the model. All DOF can have large motions without losing accuracy
of the solution except for blades and the tower DOF. FAST can also find a linear state-space representation of the
turbine at a specified operating point that can be used to design a control system using linear control theory. Since
2006, Jonkman et al.5, 8 have modified the standard version of FAST to include the necessary modules to run floating
offshore simulations. The modified FAST now includes a hydrodynamics module and a quasi-static mooring
module; the structure is shown in Fig. 2. The quasi-static mooring module solves the equations by considering the
system to be in static equilibrium at every computation time to calculate the forces in the mooring lines. FAST can
be used to interface with MATLAB/Simulink to provide the nonlinear equations of motions to be solved by the
Simulink solver.
MATLAB/Simulink is a very powerful commercial package that can be used to model systems and design
controllers of almost any type. The simulations will use the high fidelity nonlinear model of FAST in Simulink

4
American Institute of Aeronautics and Astronautics
Figure 2. Modules of offshore version of FAST13

through the FAST-Simulink interface. Controllers can then be developed in MATLAB/Simulink and tested on the
nonlinear model. Simulink was chosen instead of FAST’s own control system simulation module mainly to have the
flexibility of any control structure and relative ease of development of controllers in Simulink. A typical controller
implementation using this interface is shown in Fig. 3.
Several performance measures will be used to quantify the performance of simulated controllers allowing for
comparison between different types of controllers. Eleven performance measures are used for comparison between
the implemented controllers; these are:
ƒ RMS Speed Error: Root Mean Square (RMS) of the rotor speed error in revolutions per minute (rpm). It
serves as an indicator of rotor speed regulation; the lower the value the better the speed regulation thus
better power regulation.
ƒ RMS Pitch Rate: RMS of the blades pitch rate in degrees per second (deg/s). This is used to indicate the
frequency of actuator usage. A high value means high blade pitch actuator usage.
ƒ Max Pitch Rate: The maximum blade pitching rate among all the turbine blades in deg/s. It is used to
indicate whether the actuator is saturated to its maximum pitching speed.
ƒ Flap DEQL: Blade flapwise bending Damage EQuivalent Load (DEQL) in kNm. DEQL is a measure of
fatigue damage where the loads is calculated at a reference frequency of 1Hz.
ƒ Tower FA DEQL: Tower Fore-Aft (FA) bending DQEL in kNm.
ƒ Tower SS DQEL: Tower Side-to-Side (SS) bending DEQL in kNm.
ƒ LSS DEQL: Low Speed Shaft (LSS) torsion DEQL in kNm.
ƒ Ptfm RMS Roll: Floating platform RMS roll in deg. This is used as a measure for the platform rotation
about its roll axis; this should be kept as low as possible to reduce tower side-to-side bending loads.
ƒ Ptfm RMS Pitch: Floating platform RMS pitch in deg. This is used as a measure for the platform
rotation about its pitch axis; this should be kept as low as possible to reduce tower fore-aft bending
loads.
ƒ Ptfm RMS Yaw: Floating platform RMS yaw in deg. This is used as a measure for the platform rotation
about its yaw axis; this should be kept as low as possible to reduce rotor yaw error that will result in
reducing the amount of captured power.
ƒ Ptfm RMS Pitch Rate: Floating platform RMS pitch rate in deg/s. This is used as a measure for the
platform rotation speed about its pitch axis; this should be kept as low as possible to reduce tower fore-
aft fatigue damage and improve power regulation.
Out of these performance measures, the main emphasis will be on the RMS Speed Error, Ptfm RMS Pitch, and
Ptfm RMS Pitch Rate because the main objectives of controllers developed are to regulate rotor speed and reduce
platform pitching. Other performance measures will be used to influence the decision when the three main measures
are roughly the same.

5
American Institute of Aeronautics and Astronautics
Figure 3. LQR controller implementation using FAST-Simulink interface

B. Model Details
As this study is a continuation of the work presented in Ref. 12, the same wind turbine and barge models will be
used. Because FAST requires detailed properties about the wind turbine, therefore the turbine is a three bladed
fictional turbine with its properties derived from several similar sized available wind turbines since such details are
not publicly available11. Table 1 lists the main wind turbine and barge details.

III. Time Invariant LQR Controller

A. Baseline Controller
Before starting to design state-space controllers for the floating turbine, the new FAST-Simulink simulator has to
be verified through recreating the controller developed by Jonkman12 and simulating a baseline controller using the
two solvers given the same wind and wave conditions. This step is necessary to show that there are no significant
differences between the Simulnik differential equation solver and FAST’s solver. The best controller developed by
Jonkman, the detuned gain scheduled PI controller, will be used as the baseline for this study and all the developed
controllers will be compared to this baseline controller in region 3. Future reference to the baseline controller refers
to the detuned gains scheduled PI controller
developed by Jonkman in Ref. 12. 8
The baseline controller consists of two Table 1. General wind turbine and barge properties
Rating 5MW
parallel controllers; generator torque
Rotor Orientation Upwind
Wind Turbine

controller and collective blade pitch


controller. The torque controller varies the Control Variable speed, variable pitch, active yaw
applied generator torque as a function of
Rotor, Hub Diameter 126m, 3m
filtered generator speed. The generator
speed, used by both controllers, is filtered Hub Height 90m
using a low pass filter with a cut-off Rated Rotor, Generator Speed 12.1rpm, 1173.7rpm
frequency of 0.25Hz. This frequency was
Width 40m
chosen to be a quarter of the blades
edgewise natural frequency to prevent the Length 40m
Barge

controller from exciting these modes12; this Height 10m


behavior was also observed by Wright 14
where he found that a PI controller with a Anchor Depth 150m

6
American Institute of Aeronautics and Astronautics
fast actuator can destabilize certain modes due to the actuation frequency. The relationship between the generator
torque and generator speed is region dependent. In region 2 (below rated wind speed), the relationship that
maximizes power capture is given by Eq. (2) where ρ, RRotor, N, ωGen, CP,Max, and λo are air density, rotor radius,
gearbox ratio, generator rotational speed, maximum power coefficient and tip speed ratio that yields the maximum
CP respectively17. In region 3 where the objective is to regulate power to the rated, the relationship is given by Eq.
(3) where PRated is the rated generator power and ηGen is the generator efficiency. Furthermore, the same region
transition logic designed for smooth transition between region 2 and 3 used by Jonkman12 was implemented.

πρRRotor 5 C P , Max 2
TGen = ω Gen = Kω Gen
2
(2)
2λ o N
2 3

PRated
TGen = (3)
η Genω Gen

The gain scheduled PI controller is of the form of Eq. (4) where θ(t) is the commanded collective blade pitch
angle, KP(θ) and KI(θ) are the scheduled proportional and integral gains respectively, and e(t) is the error signal. For
details on how the gains were calculated, please refer to Ref. 12. Full DOF simulations using MATLAB/Simulink
and FAST are shown in Fig. 4 given the same wind and wave conditions of 18m/s stochastic wind and irregular sea
state with significant wave height of 3.25m and peak spectral period of 9.7s; results are almost exact. This verifies
that the Simulink implementation of the floating turbine model is correct. Thus, state-space comparisons can be
made with the baseline controller.

t
θ (t ) = K P (θ )e(t ) + K I (θ )∫ e(τ )dτ where e(t ) = ωGen − ω Gen, Rated (4)
0

B. Controller Design
A linear plant can be represented by a linear state-space system given by Eq. (5) where x and u are the state and
control input vectors respectively, A and B are the state and actuator gain matrices respectively. To control the
system, we form the feedback law given by Eq. (6), where KLQR is the Full State Feedback (FSFB) gain. With FSFB
controller we assume that all the states can be directly measured; hence, there is no need to design a state estimator
for reconstruction of system states based on measurements.

x& = A x + Bu (5)

u = − K LQR x (6)

To design a LQR controller, certain conditions have to be met. Once these conditions are met, the optimal FSFB
gain, KLQR, can be calculated from Eq. (7) where P is the solution of the Riccati equation given by Eq. (8). This gain
minimizes the quadratic cost function J given by Eq. (1). The absolute values of the Q and R matrices are not as
important as the relative difference between the two; that relative difference is the trade off between state regulation
(more weighting on Q relative to R) and actuator usage. Furthermore, the relative weighting inside each matrix
influences the importance of which states to emphasize its regulation in the case of the Q matrix and which actuator
to penalize using in case of the R matrix.

K LQR = R −1 B T P (7)

AT P + PA − PBR −1 B T P = −Q (8)

7
American Institute of Aeronautics and Astronautics
15

Rotor Speed, rpm


10

FAST Simulink
5
0 50 100 150 200 250 300

5
Platform Pitch, deg

-5
0 50 100 150 200 250 300

30
Blade Pitch, deg

20

10

0
0 50 100 150 200 250 300

50
Torque, kNm

45

40

35
0 50 100 150 200 250 300
Time, sec
Figure 4. Comparison between MATLAB/Simulink and FAST simulations

To apply this controller to the floating wind turbine, a linear model was obtained using FAST linearization about
a specified operating point using collective blade pitch; the linearization was carried out at a wind speed of 18m/s –
a point well into region 3 – and no incident waves. The model was linearized using collective pitch with two active
DOF; rotor rotational DOF and platform pitch DOF; these DOF were chosen because the controller objectives are to
regulate rotor speed and platform pitch. The linearization gives rise to 4 states; platform pitch φ (rad), rotor azimuth
ψ (rad), and their respective first derivatives in rad/s. The azimuth-averaged linearized model is given by Eq. (9)
where the state equation becomes perturbations about the linearization point with xop and uop in radians are the
equilibrium states and blades collective pitch angle respectively. Units of AAvg and BAvg matrices are: s-1 and s-2.
B

Δ x& = AAvg Δ x + BAvg Δu


op
x = Δx + x (9)
u = Δu + u op
With

8
American Institute of Aeronautics and Astronautics
⎡Δϕ ⎤ ⎡ 0 0 1 0 ⎤ ⎡ 0 ⎤
⎢Δψ ⎥ ⎢ ⎥ ⎢ ⎥
Δx = ⎢ ⎥, AAvg = ⎢ 0 0 0 1 ⎥, BAvg = ⎢ 0 ⎥
⎢Δϕ& ⎥ ⎢ − 0.2963 0 − 0.1164 − 0.0088⎥ ⎢− 0.0217⎥
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎣Δψ& ⎦ ⎣− 0.0556 0 − 2.9202 − 0.2406⎦ ⎣ − 0.3924⎦
⎡0.018⎤
⎢3.137 ⎥
x =⎢ ⎥, u op = 0.259
op

⎢0 ⎥
⎢ ⎥
⎣1.267 ⎦

With the linearized state-space model available, the optimal FSFB gain can be calculated using a MATLAB
routine that checks the LQR assumptions and solves the Riccati equation. The routine scales the B matrix to avoid
large variations between different actuators. This is useful when the generator torque is included in the state-space
controller because of their different effect on the linearized states that could results in B matrix with columns orders
of magnitudes different. For this controller, the generator torque is held constant. Scaling the B matrix was done by
normalizing every column of the matrix; the new scaled B matrix becomes [0 0 -0.055 -1]T. The scaling factor is
then multiplied by the corresponding column of the commanded actuator signal Δu. The Q and R matrices are
specified in that routine. The R matrix is kept as an identity matrix while the Q matrix is changed to find the best
controller performance. The LQR implementation using Simulink is shown in Fig. 3 with results discussed in the
next section.
The generator torque is held constant instead of using the baseline torque controller method because we carried
out a small study (not included here) that concluded with a constant generator torque the controller the rotor speed
damping is increased resulting in better speed regulation and with less DEQL, however, variable torque or constant
power has better power regulation but would require the pitch controller to compensate for reduced rotor speed
damping.

C. Simulation Results
Two types of simulations were carried out. First, the designed controller based on a 2DOF model is tested with
the nonlinear 2DOF model to see if the system remains stable and to do initial controller tuning. Then, the same
controller is then simulated with all DOF enabled in the FAST model. The incident wind and wave conditions
(18m/s stochastic wind and irregular sea state with significant wave height of 3.25m and peak spectral period of
9.7s) used in simulations are shown in Fig. 5.
25
Wind Speed, m/s

20

15

10
0 50 100 150 200 250 300
4
Wave Elevation, m

-2
0 50 100 150 200 250 300
Tim e, sec
Figure 5. Hub-height wind speed and incident wave elevation used in simulations.

9
American Institute of Aeronautics and Astronautics
Table 2. Comparison between performance of baseline and two collective pitch
LQR controllers for 2DOF simulations.
Performance Index Baseline LQR1 % Change LQR2 % Change

RMS Speed Error (rpm) 1.21 1.08 -11% 1.20 -1%


RMS Pitch Rate (deg/s) 0.77 1.38 78% 1.76 127%
Max Pitch Rate (deg/s) 1.94 4.37 125% 5.39 178%
Flap DEQL (kNm) 11,950 12,085 1% 13,657 14%
Tower FA DEQL (kNm) 46,621 33,697 -28% 31,114 -33%
Tower SS DEQL (kNm) 2,428 2,533 4% 2,450 1%
LSS DEQL (kNm) 657 503 -23% 565 -14%
Ptfm RMS Roll (deg) 0.00 0.00 NA 0.00 NA
Ptfm RMS Pitch (deg) 2.45 1.86 -24% 1.79 -27%
Ptfm RMS Yaw (deg) 0.00 0.00 NA 0.00 NA
Ptfm RMS Pitch Rate (deg/s) 1.19 0.91 -23% 0.89 -25%

1. Two Degrees of Freedom Simulations


Since Jonkman’s results are for all DOF enabled in the simulation, the first simulation to be carried out was the
baseline controller with 2DOF enabled in the simulation. Then, several LQR controllers were simulated and the best
two controllers’ results are listed in Table 2 along with the baseline controller. The first controller, LQR1, has Q =
daig(0.3, 0.3, 300, 0.3) – these are the diagonal elements of the Q matrix and all other elements are zero. The second
controller, LQR2, has Q = diag(0.5, 0.1, 100, 0.5). All simulations were given the same wind and wave conditions.
The % Change column is calculated relative to the baseline controller. Rows highlighted in yellow are the main
performance indices that are used to judge the performance of controllers.
Both LQR controllers perform better than the baseline controller; all the performance measures are reduced with
the exception of those related to actuator usage and Flap DEQL. A percentage change of within 5% is considered
negligible due to statistical noise and variation. Increased actuator usage was expected due to the fact that LQR
controllers have multiple objectives and access to only one actuator; collective blade pitch. The more aggressive
blade pitch actuation in LQR2 increases cyclic loading on the blades in the flapwise direction.
Comparing the LQR controllers with each other, it can be seen that LQR2 regulates the platform pitch and
velocity slightly better than LQR1 but reduced speed regulation. Furthermore, LQR1 has less actuator usage which
reduces the loads on the blades in the flapwise direction as shown in Table 2. Therefore, with better speed
regulation, less actuator usage and comparable platform pitching, LQR1 is the better controller for the 2DOF
simulation.
2. Full Degrees of Freedom Simulations
The same collective pitch LQR controllers that were found among the best controllers in 2DOF simulations are
now simulated with all DOF enabled in the simulation. The aim here is to see if the controllers maintain system

Table 3. Comparison between performance of baseline and two collective pitch


LQR controllers for all DOF simulations.
Performance Index Baseline LQR1 % Change LQR2 % Change

RMS Speed Error (rpm) 1.08 0.85 -21% 0.88 -18%


RMS Pitch Rate (deg/s) 0.69 1.11 60% 0.72 4%
Max Pitch Rate (deg/s) 1.98 4.18 111% 2.76 39%
Flap DEQL (kNm) 9,829 9,620 -2% 8,962 -9%
Tower FA DEQL (kNm) 37,128 30,001 -19% 32,057 -14%
Tower SS DEQL (kNm) 14,068 16,740 19% 16,446 17%
LSS DEQL (kNm) 1,227 708 -42% 682 -44%
Ptfm RMS Roll (deg) 0.36 0.28 -21% 0.32 -12%
Ptfm RMS Pitch (deg) 1.64 1.35 -18% 1.45 -11%

Ptfm RMS Yaw (deg) 2.79 2.72 -3% 2.88 3%


Ptfm RMS Pitch Rate (deg/s) 0.69 0.53 -23% 0.61 -11%

10
American Institute of Aeronautics and Astronautics
stability while improve the performance of the system relative to the baseline controller.
Results are listed in Table 3 where it can be seen that once again both LQR controllers perform better than the
baseline gain scheduled PI controller. Similarly to 2DOF simulations, LQR1 has better speed regulation and less
actuator usage but has better platform pitch regulation when compared with LQR2.
Similarly to 2DOF simulations, LQR1 is the best controller which has improvement across all performance
indices with the exception of blade pitch actuator usage and Tower SS DEQL. The increased side-to-side tower
loading could be due to the controller trying to regulate the platform pitch through increased collective blade
pitching; as the turbine pitches forward, the controller commands the blades to increase aerodynamic thrust and
consequently the aerodynamic torque. For a three bladed wind turbine, the generated aerodynamic torque is
balanced given a uniform incident wind; but, due to wind shear, a blade at the top generates more torque and thus
creates a net sideways force.

16
Rotor Speed, rpm

14

12

10
Baseline LQR1 Rated Speed
8
0 50 100 150 200 250 300

4
Platform Pitch, deg

-2

-4
0 50 100 150 200 250 300

30
Blade Pitch, deg

20

10

0
0 50 100 150 200 250 300

50
Gen Torque, kNm

45

40

35
0 50 100 150 200 250 300
Tim e, sec
Figure 6. Full DOF simulation output showing baseline controller and LQR1.

11
American Institute of Aeronautics and Astronautics
Figure 6 shows the simulation output of the baseline
and LQR1 controllers. Four signals are plotted; rotor speed
and platform pitch which correspond to the controller
objectives and collective blade pitch and applied generator
torque which shows actuator usage. The figure shows
improvement of rotor speed and platform pitch regulation
as well as increased blade pitching.

D. Discussion of Collective Pitch Control


We have shown that LQR controller performs better
than baseline PI controller, but the questions is how and
why is it performing better? The first obvious reason is that
the baseline controller considered in this study is designed
to regulate rotor speed only whereas the LQR controllers
are designed to regulate rotor speed and platform pitch.
However, the controller with tower top feedback (TTF)
designed by Jonkman (see Section I) is essentially similar
to the LQR controller, but its performance was even worse
than the baseline controller considered in this study. Since
TTF and LQR controllers use collective blade pitch, the
platform pitch restoring mechanism is the same for both
controllers. The mechanism works by reducing blade pitch
as the turbine is pitching forward thus increasing thrust and
torque as shown in Fig. 7. However, this increase in thrust
also increases aerodynamic torque hence accelerating the
rotor. Therefore, the two objectives of speed and platform
pitch regulation are competing for blade pitch. This is what
essentially Jonkman found when he added the tower top
feedback control loop; the two independent controllers
were competing for blade pitch usage. The LQR controller
is better because of tuning. The Q matrix was chosen to put
a large emphasis on regulating platform pitch velocity state
to 0° which essentially allows the platform pitch regulator
to dominate over the speed regulator part of the controller.
If the tower feedback loop was given larger gains, similar
performance may be achieved by the gain scheduled PI Figure 7. Platform pitch restoring mechanism
controller. The use of constant torque instead of constant using collective blade pitch.
power is also responsible for improved speed regulation by
the LQR controller due to increased rotor speed damping.
Although difference in performance may be attributed to controller tuning, LQR controller offers the flexibility
to include more design objectives, accommodate more actuators, and multiple sensors. State estimation,
reconstructing system states given limited measurements, is outside the scope of this paper. If a state estimator was
added, the performance of the FSFB controller will degrade. However, for this controller the performance
degradation is not expected to be severe because most of the controller states can be easily measured and the state
estimates is expected to be close to the actual.
Another issue worth noting is that platform motion is not only dominated by pitch. Other motions are as large as
platform pitch if not even larger as shown in Fig. 8. However, platform pitch has significant impact on power and
speed regulation.

IV. Individual Blade Pitch Controller Design

A. Controller Design
Because both TTF and collective pitch LQR controllers use the same mechanism to regulate platform pitch, the
difference in performance comes down to tuning; the LQR controller is an optimal way of calculating the gains to
meet certain objectives. Therefore, we now investigate whether Individual Blade Pitch (IBP) control uses a different
mechanism to restore platform pitch.

12
American Institute of Aeronautics and Astronautics
Platform Linear Displacem ents Platform Angular Displacem ents
35 8
Surge Sw ay Heave Roll Pitch Yaw
30 6

25
4
Displacement, m

20

Angle, deg
2
15
0
10
-2
5

0 -4

-5 -6
0 50 100 150 200 250 300 0 50 100 150 200 250 300
Tim e, sec Tim e, sec
Figure 8. Platform motion with all degrees of freedom active in simulation.

There are many ways to achieve IBP control depending on the control objectives. Bossanyi18 implemented IBP
control using PI controllers to mitigate blade loads. He used a direct and quadrature (d-q) axis representation to be
able to allow for MIMO control using PI controllers. Wright14 achieved IBP control using disturbance
accommodating controller, but he used an internal model for the wind disturbance and effect of wind shear to drive
the individual blade pitching. Another form of IBP is to use periodic control. Periodic control allows the controller
gains to change depending on the rotor azimuth position and control objectives15, 19. Periodic control will be used to
implement IBP control to regulate rotor speed and platform pitch.
To design a periodic LQR controller, first a periodic state-space model must be extracted from FAST. During
linearization, FAST generates the periodic state-space matrices required for this controller design for a specified
number of azimuth positions16; usually these matrices are averaged to design constant gain controllers (the same
method we used to design LQR controller in Section III).
The linearized periodic state-space model is given by Eq. (10) where A(t) and B(t) are the periodic state and
control gain matrices respectively. The periodic feedback control law, Eq. (11), is similar to the standard state-space
controller except for the gain being periodic; the negative feedback sign is incorporated into the periodic gain G(t).

Δ x& = A(t )Δ x + B(t )Δu (10)

Δu = G (t )Δ x (11)

The optimal periodic gain, periodic LQR gain, is given by Eq. (12) where P(t) is the solution of periodic Riccati
equation given by Eq. (13)19.

G (t ) = − R −1 B (t ) P (t )
T
(12)

P& (t ) + A(t ) P (t ) + P (t )A(t ) − P (t )B (t )R −1 B(t ) P (t ) = −Q


T T

P (t final ) = Pfinal
(13)

Solving the periodic Riccati equation using MATLAB allows us to calculate the optimal periodic gains. The
periodic gain for blade 1 corresponding to platform pitch velocity, 3rd state, is shown in Fig. 9 along with the
equivalent constant LQR gain found using the averaged state-space model; both were calculated from the same Q
matrix. The zero azimuth position is at the 12 o’clock position on the rotor with angle increasing counterclockwise
when looking downwind; initially blade 1 is at the 0 azimuth position. Interestingly, the constant gain derived from
the averaged model is not the average of the periodic gain. Furthermore, the periodic gain changes sign twice as the

13
American Institute of Aeronautics and Astronautics
blade goes through a full rotor revolution. The figure 15
does not show the gains for the other two remaining Periodic
blades as the gain is the same but shifted 120° and 240° 10 Constant
out of phase for blades 2 and 3 respectively. The
significance of the sign change will be explained in the
5
next section.

Gains
B. Physical Principles behind IBP Control 0
In Section III D the mechanism by which collective
pitch controller generate the restoring pitch moment was -5
discussed and illustrated in Fig. 7. To see whether IBP
control uses the same mechanism or a different one, we
-10
look at clues in the periodic gain matrix; in particular, the 0 60 120 180 240 300 360
gain corresponding to the platform pitch velocity shown Azim uth, deg
in Fig. 9.
First, we note that the gain is negative for azimuth Figure 9. Blade 1 periodic and constant LQR gains
angles between 105° and 245°; these angles correspond to for the platform pitch velocity state (state 3)
the blade lying in the lower half of the rotor. Second, by
inspecting Eq. (11), we can determine the direction of
blade pitch actuation given the change in the system
states. Therefore, given a negative platform pitch
velocity, shown in Fig. 10, blades at the top with a
positive controller gain will be commanded to reduce
blade pitch thus increasing thrust. Blades at the bottom
with a negative controller gain will be commanded to
increase blade pitch and hence reducing thrust. This
asymmetric aerodynamic loading will generate a positive
restoring pitching moment in addition to the restoring
moment generated by the mean thrust load (as with
collective pitch) as illustrated in Fig. 10. Conversely, if
the platform pitch velocity is positive, then the controller
will command blades at the top to increase blade pitch
thus reducing thrust while it will command blades at the
bottom to reduce blade pitch increasing thrust; this
differential thrust creates a platform pitch restoring
moment.
Looking further at the periodic gain shown in Fig. 9,
we can see that it is not symmetrical around zero. Almost
two thirds of the gain across rotor azimuth is positive and
with higher positive values most of which is when the
blade is close to the top position. More positive gain
value means increased blade pitching, which indicates
that the controller is making use of the combined effect of
increased moment arm and wind shear.
The periodic gain that corresponds to the platform
pitch state, 1st state, is not shown but has the same shape
as in Fig. 9 when this state is emphasized in the Q matrix.
That is, the gain that corresponds to the platform pitch
angle is very sensitive to controller tuning and Q matrix
weightings. The specific shape of the periodic gain, when
it is positive when the blade is at the top and negative or
smaller at the bottom, drives the controller to regulate
platform pitch. When less weighting is put for this state,
the periodic gain would not retain this shape; fortunately,
the gain that corresponds to the 3rd state retains that shape Figure 10.Platform pitch restoring mechanism using
and ensures platform pitch velocity regulation. individual blade pitch.

14
American Institute of Aeronautics and Astronautics
Table 4. Comparison between the performance of collective vs. individual blade pitch
LQR controllers for 2 DOF simulations.
Incident Wave 1 Incident Wave 2
Performance Index LQR2 PLQR % Change LQR2 PLQR % Change
RMS Speed Error (rpm) 1.04 1.00 -4% 0.99 0.96 -3%
RMS Pitch Rate (deg/s) 0.86 2.57 199% 0.85 2.58 205%
Max Pitch Rate (deg/s) 2.71 8.00 195% 5.20 8.00 54%
Flap DEQL (kNm) 11,055 14,539 32% 10,256 13,843 35%
Tower FA DEQL (kNm) 37,578 36,558 -3% 38,989 38,057 -2%
Tower SS DEQL (kNm) 2,468 3,476 41% 2,479 3,467 40%
LSS DEQL (kNm) 445 418 -6% 451 420 -7%
Ptfm RMS Roll (deg) 0.00 0.00 NA 0.00 0.00 NA
Ptfm RMS Pitch (deg) 2.05 1.90 -7% 1.98 1.83 -8%
Ptfm RMS Yaw (deg) 0.00 0.00 NA 0.00 0.00 NA
Ptfm RMS Pitch Rate (deg/s) 0.89 0.90 1% 1.03 0.90 -13%

Furthermore, the negative sign is not necessary to ensure platform pitch regulation; the mechanism works by
creating a differential thrust, the negative sign ensures that this differential thrust is maximized.

C. Simulation Results
In Section III we have shown that collective pitch LQR controller is able to improve platform pitch regulation
while maintaining or improving rotor speed regulation over the baseline controller and thus in turn better than the
TTF controller. But, this improvement is thought to be due to better tuning of the controller since both controllers,
gain scheduled PI with TTF control loop and LQR, use the same platform pitch restoring mechanism.
Several simulations were run to determine whether IBP controller which uses a different platform pitch restoring
mechanism is able to further improve the system performance. To eliminate controller tuning from affecting
simulation results, both controllers were designed using the same Q and R matrices. This is justified since both
controller are LQR and the only difference is that periodic controller changes its gain according to changes in the
A(t) and B(t) matrices while the collective pitch controller uses the azimuth averaged AAvg and BAvg matrices. B

Furthermore, two situations were simulated each with different incident waves; the only difference is the change of
the random number seed used to generate incident waves (all other wave properties are kept the same). Incident
Wave 1 is the same wave profile used in previous simulations (shown in Fig. 5) and Incident Wave 2 is the new
incident wave profile. All simulations used the same turbulent wind file since the barge motion is most sensitive to
incident waves rather than incident wind13. All simulations were run with only 2 DOF active to eliminate the effects
of any coupling or instability on results.
Initially, the periodic LQR controller was designed with the same Q matrix for LQR1 for comparison. However,
simulation shows that the periodic LQR, although has better performance than LQR1, saturates the blade pitch
actuators frequently. This is undesirable as blade pitch saturation degrades controller performance and invalidates
the results for comparison. Therefore, periodic LQR is now designed and compared with LQR2. Simulation results
are listed in Table 4 where LQR2 refers to the collective pitch controller using LQR2 Q and R matrices and PLQR
refers to the periodic or IBP controller using the same matrices as LQR2. The % Change column calculates the
percentage difference between the IBP controller and collective pitch controller of the corresponding performance
index.
Results show that periodic control using IBP achieves similar rotor speed regulation while improving platform
pitch regulation adequately in both situations. Platform pitch rate is similar for the first incident wave profile but it is
better regulated by the periodic controller in the second incident wave profile. The increased actuation of the blades
is also reflected in the results; the maximum pitch rate is actually the saturation limit for the actuator. However,
looking at the blades pitching rates (not shown here), the blades are only saturated at two instances both of which are
less than 1 second in duration; this is acceptable as its impact on the performance of the system is negligible if any.
Also as expected, the blades flapwise DEQL is higher for the periodic controller while tower fore-aft DEQL is
similar in both situations. However, tower side-to-side DEQL has increased considerably which indicates that the
periodic controller has an undesired influence on system modes that are not included in the controller design. This

15
American Institute of Aeronautics and Astronautics
25

20
Blade Pitch, deg

15

10

5 Collective
IBP, Blade 1
IBP, Blade 2
IBP, Blade 3
0
0 5 10 15 20 25 30 35 40 45 50
Time ,sec
Figure 11.Commanded blade pitch for collective and IBP controllers.

could cause instability with the fully coupled wind turbine. Full DOF simulation and analysis for the periodic
controller is outside the scope of this paper.

D. Discussion of Individual Blade Pitch control


The flapwise DEQL for the blades has increased considerably. This is attributed to increased blade pitch
actuation due to the periodic controller. Figure 11 shows a plot of commanded blade pitches for collective and IBP
controllers. The periodicity of the IBP controller is obvious with each blade 120º out of phase, but in addition to
increased actuation due to periodicity, the actual magnitude is larger. This explains the significant increase in the
blades flapwise DEQL. The figure also shows that the mean of all blades follows the same trajectory as the
commanded collective blade pitch; this confirms that IPB controller is generating asymmetric aerodynamic thrust
loads through the periodic actuation of each blade as well as using the mean thrust through the mean blade pitch (as
discussed in Section IV B).
Initially, the cause of the significant increase in tower side-to-side DEQL is not obvious. But looking closely at
the some of the turbine properties through linearization, a relationship was found. The increased tower side-to-side
loading is due to edgewise force applied by the blades. Therefore, to look for interaction, the relationship between
the effect of the blades on platform roll is studied. Interestingly, the blades affect the platform roll the same way it
affects platform pitch and the effect is in phase. This means that when the controller generates the restoring pitch
moment it also induces a rolling moment. However, because the simulations were run with only 2 DOF active, that
effect is reflected through increased tower side-to-side loads.
The platform pitch restoring mechanism by which IBP use was proven to produce better performance over the
collective pitch LQR controller given the same Q matrix. However, IBP control has undesired effects on tower side-
to-side bending loads. But adding other design objectives such as loads reduction is easy with state-space controller.
Furthermore, Stol et al.15 and Bossanyi20 showed that individual blade pitch control is capable of reducing loads
better than collective pitch controllers.

V. Conclusions
For the work presented in this paper, we conclude that:
ƒ A collective blade pitch controller using a linear quadratic regulator (LQR) gives better performance in
terms of rotor speed regulation as well as platform pitch regulation over a gain scheduled PI controller
designed for rotor speed regulation only. The designed LQR controller was able to reduce the root mean

16
American Institute of Aeronautics and Astronautics
square of the rotor speed error by 21%, reduce platform pitch oscillations by 18% and platform pitch
rate by 23%.
ƒ This improved performance is due to the use to constant torque control instead of constant power
control for speed regulation and possibly better controller tuning. If an additional control loop was
added to the gain scheduled PI controller and given more weighting, similar performance to the
collective pitch LQR controller may be achieved because both controller use the same platform pitch
restoring mechanism.
ƒ For collective pith controllers, the platform pitch restoring mechanism is achieved through changing the
uniform thrust on the rotor. However, for an individual blade pitch controller the mechanism used is a
combination of changing the symmetric thrust as well as inducing an asymmetric rotor load thus
creating a pitch restoring moment.
ƒ Initial simulations using a low fidelity model indicate that individual blade pitch controller is capable of
further improving the performance obtained by the collective pitch LQR controller. However, this
improved performance using the differential thrust mechanism is also inducing a rolling moment that
could destabilize the system if not taken into account; further analysis is required to quantify the
performance of using individual blade pitch controller on floating offshore wind turbines.

Future Work
Future work on this topic will include closed loop stability analysis for the designed controllers. The authors will
be investigating the performance of using individual blade pitch controllers with more degrees of freedom enabled
in the FAST model, a higher fidelity model. Furthermore, more control objectives will be included in the design of
the controllers such as adding damping to the drive train and reducing tower fore-aft loads. Once a satisfactory
controller performance is achieved, the effects of adding a state estimator to the system will be investigated.

Acknowledgments
H. Namik would like to thank The University of Auckland for student funding through a doctoral scholarship.
He would also like to thank the Wind Engineering Group at the Department of Mechanical Engineering for
providing feedback during the year.

References
1
Mathew, S., Wind Energy : Fundamentals, Resource Analysis and Economics, Springer, Berlin, 2006.
2
Shikha, T.S. Bhatti, and D.P. Kothari, "Aspects of Technological Development of Wind Turbines," Journal of Energy
Engineering, Vol. 129, No. 3, 2003, pp. 81-95.
3
Brennan, S., Offshore Wind Energy Resources, in Workshop on Deep Water Offshore Wind Energy Systems. 2003,
Department of Energy, The Office of Energy Efficiency & Renewable Energy, and National Renewable Energy Laboratory.
4
Nielsen, F.G., T.D. Hanson, and B. Skaare, "Integrated Dynamic Analysis of Floating Offshore Wind Turbines."
Proceedings of the25th International Conference on Offshore Mechanics and Arctic Engineering - OMAE, Hamburg, 2006.
5
Jonkman, J.M. and P.D. Sclavounos, "Development of Fully Coupled Aeroelastic and Hydrodynamic Models for Offshore
Wind Turbines," Proceedings of the 44th AIAA Aerospace Sciences Meeting and Exhibit, 2006, pp. 9-12.
6
"Largest and deepest offshore wind turbine installed," Refocus, Vol. 7, No. 5, 2006, pp. 14.
7
Musial, W., S. Butterfield, and B. Ram, "Energy from Offshore Wind." Offshore Technology Conference, Houston, Texas,
USA, 2006.
8
Jonkman, J.M. and M.L. Buhl Jr., "Development and Verification of a Fully Coupled Simulator for Offshore Wind
Turbines." 45th AIAA Aerospace Sciences Meeting and Exhibit, Reno, Nevada, 2007, pp. 28-53.
9
Vijfhuizen, W.M.J.M., "Design of Wind and Wave Power Barge," M.Sc. Thesis, Department of Naval Architecture and
Marine Engineering, Universities of Glasgow and Strathclyde, Glasgow, Scotland, 2006.
10
Wayman, E.N., "Coupled Dynamics and Economic Analysis of Floating Wind Turbine Systems," M.Sc. Thesis, Dept. of
Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts, 2006.
11
Jonkman, J., S. Butterfield, W. Musial, and G. Scott, "Definition of a 5-MW Reference Wind Turbine for Offshore System
Development," National Renewable Energy Laboratory TP-500-38060, Golden, CO, 2007 (to be published).
12
Jonkman, J.M., "Dynamics Modeling and Loads Analysis of an Offshore Floating Wind Turbine," Ph.D. Thesis,
Department of Aerospace Engineering Sciences, University of Colorado, Boulder, Colorado, 2007.
13
Jonkman, J.M. and M.L. Buhl Jr, "Loads Analysis of a Floating Offshore Wind Turbine Using Fully Coupled Simulation."
WindPower 2007 Conference & Exhibition, Los Angeles, California, USA, 2007.
14
Wright, A.D., "Modern Control Design for Flexible Wind Turbines," National Renewable Energy Lab NREL/TP-500-
35816, Golden, CO., 2004.

17
American Institute of Aeronautics and Astronautics
15
Stol, K.A., W. Zhao, and A.D. Wright, "Individual blade pitch control for the Controls Advanced Research Turbine
(CART)," Journal of Solar Energy Engineering, Transactions of the ASME, Vol. 22, 2006, pp. 16478-16488.
16
Jonkman, J. and M.L. Buhl Jr, "FAST User's Guide," National Renewable Energy Laboratory NREL/EL-500-38230,
Golden, Colorado, 2006.
17
Bossanyi, E.A., "The Design of Closed Loop Controllers for Wind Turbines," Wind Energy, Vol. 3, No. 3, 2000, pp. 149-
163.
18
Bossanyi, E.A., "Individual Blade Pitch Control for Load Reduction," Wind Energy, Vol. 6, No. 2, 2003, pp. 119-128.
19
Stol, K., "Dynamics Modeling and Periodic Control of Horizontal-Axis Wind Turbines," PhD Thesis, Department of
Aerospace Engineering Sciences, University of Colorado, Boulder, Colorado, 2001.
20
Bossanyi, E.A., "Wind Turbine Control for Load Reduction," Wind Energy, Vol. 6, No. 3, 2003, pp. 229-244.

18
American Institute of Aeronautics and Astronautics

View publication stats

You might also like