Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

594 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 24, NO.

2, MARCH 2016

Automatic Retraction and Full-Cycle Operation for


a Class of Airborne Wind Energy Generators
Aldo U. Zgraggen, Lorenzo Fagiano, Member, IEEE, and Manfred Morari, Fellow, IEEE

Abstract— Airborne wind energy systems aim to harvest The wing’s path can be influenced by means of different
the power of winds blowing at altitudes higher than what technical solutions, which typically give rise to a steering
conventional wind turbines reach. They employ a tethered flying input corresponding to a change of the roll angle of the
structure, usually a wing, and exploit the aerodynamic lift to
produce electrical power. In the case of ground-based systems, wing. Assuming a straight tether, the path of the wing is
where the traction force on the tether is used to drive a generator restricted to a spherical surface with a radius equal to the
on the ground, a two-phase power cycle is carried out: one phase tether length, confined by the ground and a vertical plane
to produce power, where the tether is reeled out under high perpendicular to the wind direction. This spherical surface
traction force, and a second phase where the tether is recoiled is commonly called wind window. Depending on the path
under lower load. The problem of controlling a tethered wing
in this second phase, the retraction phase, is addressed here, by flown by the wing, a higher or lower traction force is expe-
proposing two possible control strategies. Theoretical analyses, rienced. During crosswind paths, a high wing speed can be
numerical simulations, and experimental results are presented achieved and thus a high traction force is exerted. On the
to show the performance of the two approaches. Finally, the other hand, if the wing is flown on the side of the wind
experimental results of complete autonomous power generation window, i.e., with the tether roughly perpendicular to the wind
cycles are reported and compared with those in first-principle
models. direction, a low wing speed results and a small traction force is
exerted.
Index Terms— Airborne wind energy (AWE), control These two different flying conditions can be exploited in
applications, control of tethered wings, high-altitude wind
power, kite power, wind energy. ground-based generation AWE systems by flying a two-phase
power cycle [5], [6]. In the first phase, called traction phase,
I. I NTRODUCTION power is produced by flying a crosswind pattern and using the
high traction force to unreel the tether from the drum. Once the
A IRBORNE wind energy (AWE) systems are an emerging
technology to harvest renewable energy from wind. Their
aim is to harness the energy contained in the strong and steady
maximum tether length has been reached, the second phase,
called retraction phase, is carried out by moving the wing on
winds beyond the altitude reached by traditional wind turbines the side of the wind window and then recoiling the lines under
(see [1], [2] for an overview). These systems consist of a low traction forces. In this way, only a fraction of the energy
ground unit (GU), a wing, and one or more tethers connecting previously produced is consumed. This approach is considered
them. by various companies and research groups [4], [7]–[18].
During power production, the wing is flown in a crosswind The automatic control of tethered wings plays a major role
pattern, i.e., roughly perpendicular to the wind flow, exceeding for the operation of this kind of system and has been studied
the speed of the wind and thus exerting high aerodynamic in [6], [15], and [19]–[26]. Several of these approaches
forces. The generators can either be placed on board the wing consider only the problem of flying crosswind trajectories
or on the ground inside the GU. On-board generation systems when energy is produced. However, for ground-based genera-
use propellers driven by the high apparent wind speed and then tion systems, also the retraction of the tether has to be done
transfer the produced power to the ground via an electrified autonomously. In [6] and [19], two controllers for the retrac-
tether (see [3]). On the other hand, ground-based generation tion phase, using nonlinear model predictive control strategies,
systems use the traction force on the cable to spin a drum have been proposed. However, these strategies might be
installed on the GU and connected to a generator (see [4]). difficult to implement and tune due to their complexity.
In this paper, we consider the latter approach. In addition, they assume quite a good knowledge of the
wind speed at the wing’s location, which is hard to obtain
Manuscript received August 28, 2014; accepted June 7, 2015. Date of in practice, and they have been tested in simulations only,
publication September 4, 2015; date of current version February 17, 2016. assuming that the model used for the control calculation
Manuscript received in final form June 27, 2015. This work was supported
by the Swiss Competence Center Energy and Mobility. Recommended by corresponds exactly to the real system.
Associate Editor L. Marconi. In this paper, we tackle the problem of autonomous retrac-
A. U. Zgraggen and M. Morari are with the Automatic Control tion phase for ground-based AWE systems by presenting
Laboratory, ETH Zurich, 8092 Zurich, Switzerland (e-mail:
zgraggen@control.ee.ethz.ch; morari@control.ee.ethz.ch). two possible control approaches, which we tested in real-world
L. Fagiano is with ABB Switzerland Ltd., Corporate Research, Baden 5405 experiments with a small-scale prototype. The first one is an
Dättwil, Switzerland (e-mail: lorenzo.fagiano@ch.abb.com). extension of the approach presented in [25] and it is based on
Color versions of one or more of the figures in this paper are available
online at http://ieeexplore.ieee.org. the notion of the velocity angle of the wing, which represents
Digital Object Identifier 10.1109/TCST.2015.2452230 its flying direction. As we will show in this paper, this notion
1063-6536 © 2015 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

Authorized licensed use limited to: Consortium - Algeria (CERIST). Downloaded on June 08,2023 at 05:20:22 UTC from IEEE Xplore. Restrictions apply.
ZGRAGGEN et al.: AUTOMATIC RETRACTION AND FULL-CYCLE OPERATION FOR A CLASS OF AWE GENERATORS 595

can be adapted such that it can also be used for feedback


control during the retraction phase, when the speed of the
wing relative to the GU is low and the original definition of the
velocity angle is not valid anymore. The resulting controller is
dependent on an estimate of the wind direction at the wing’s
location. We will show that with this approach the wing can
be stabilized at the border of the wind window during the
retraction phase.
Since an estimate of the wind direction at the wing’s loca-
tion is not straightforward to obtain, we propose an alternative
approach by directly controlling the elevation of the wing.
To do so, we derive a new model relating the steering input
to the vertical acceleration of the wing, and we use such a
model for control design. In addition, this control system is
able to stabilize the wing’s trajectory at a constant elevation
angle and it exploits only directly measurable variables, hence
resulting in a more reliable and robust solution with respect
Fig. 1. Front view of the Swiss Kite Power prototype built at Fachhochschule
to the previous one. The considerations above are supported Nordwestschweiz. The two steering lines, left (red) and right (blue), are wound
by simulation results used to compare the two approaches. around drums connected via a belt drive to motors mounted below the drums.
Real-world experiments are then presented and analyzed to The center line (yellow) wound around the main drum is behind the two other
drums and is partly visible below the left steering line’s drum. All three lines
evaluate both control strategies. There exists evidence in [1] are guided separately via pulleys to the lead-out sheaves, visible at the top.
that other groups and companies have achieved autonomous On the lead-out sheave of the main line, a line angle sensor is mounted.
power cycles; however, there are no publications explaining A wind sensor, mounted roughly 5 m above the ground, is visible in the
background.
the employed control strategy. By achieving an autonomous
retraction phase, we have been able to test fully automatic
power cycles in experiments, whose results we compare here
wings or power kites. In three-line systems, the line wound
with the well-known equations [5] that lie at the very founda-
around the middle drum, called main line, is connected to
tions of the concept of AWE, showing a promising matching
the leading edge of the wing and sustains the main portion
between mathematical models and real-world data.
of the traction force. The lines on the other two drums are
Note that in this paper, we considered as manipulated inputs
called steering lines and are connected to the left and right
only the steering deviation and the force applied by the ground
wing tips. These two lines are used to influence the wing’s
generators/motors on the lines, i.e., we did not consider an
trajectory. By changing the difference δ between the length of
active control of the pitch of the wing. The latter can be
the two steering lines, the desired steering deviation can be
added on top of the presented control strategies to optimize
issued. A shorter left line induces a counter-clockwise turn of
the power production, as proposed, for example, in [27].
the wing as seen from the GU, and vice versa. The system
In addition, we only considered a retraction phase flown on
has a total rated power of 20 kW; the generator of the middle
the side of the wind window and did not consider a glide
drum has a power rating of 10 kW and each of the motors
maneuver directly downwind since we employ flexible wings.
connected to the drums of the steering lines has a power
For a glide maneuver, which indeed can be beneficial for rigid
rating of 5 kW. The system is operated with tether lengths up
wings, the angle of attack has to be kept at low values. This is
to 200 m. We first recall a dynamical model of the described
not possible for flexible wings since the angle of attack cannot
system, followed by the definition of the velocity angle γ,
be estimated and controlled with the required precision, and
which acts as one of the main feedback variables during the
the wing might even front-stall and collapse with a consequent
traction phase (for the details on the controller employed in
loss of controllability and eventual crash.
this phase, we refer the reader to [25]).
The paper is structured as follows. In Section II, we describe
the considered system and the models we use. In Section III,
we introduce the two different control approaches for the A. Model Equations
retraction phase and discuss the tether reeling scheme. The dynamical model we consider has been widely used in
In Section IV, the simulation and experimental results are previous works (see [6] and references therein). We will recall
presented and discussed for both control approaches. The the model equations shortly, following the same notation as
conclusions and future developments are given in Section V. in [25] and additionally include a further degree of freedom to
account for the reeling capabilities of the considered prototype.
II. S YSTEM D ESCRIPTION We will denote vector-valued variables in bold, e.g., G p(t),
The system under consideration is related to the Swiss Kite where the subscript letter in front of vectors denote the
Power prototype [7] (see Fig. 1). It is an AWE system featuring reference system considered to express the vector components
ground-based steering actuators with the generators placed and t denotes the time dependence.
inside the GU. It has three drums with a motor connected An inertial frame centered at the GU is denoted by
.
to each one, and it can be used with one-, two-, and three-line G = (ex , e y , ez ), where unit vectors are denoted by e with

Authorized licensed use limited to: Consortium - Algeria (CERIST). Downloaded on June 08,2023 at 05:20:22 UTC from IEEE Xplore. Restrictions apply.
596 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 24, NO. 2, MARCH 2016

with e N . The system L is a function of the wing’s position


only. The transformation matrix to express the vectors in the
local frame L from the inertial frame G is denoted by A LG
(e.g., L p(t) = A LG G p(t))
⎛ ⎞
− cos (ϕ) sin (ϑ) − sin (ϕ) sin (ϑ) cos (ϑ)
A LG = ⎝ − sin (ϕ) cos (ϕ) 0 ⎠.
− cos (ϕ) cos (ϑ) − sin (ϕ) cos (ϑ) − sin (ϑ)
(2)

From the differentiation of (1) and using the rotation


matrix (2), we obtain the velocity vector of the wing in local
coordinates L with respect to the GU
⎛ ⎞
r (t)ϑ̇(t)
⎜ ⎟
L v P (t) = ⎝r cos (ϑ(t))ϕ̇(t)⎠. (3)
Fig. 2. Wing’s position p (black dot) is shown on a figure-eight crosswind
path together with the local coordinate frame L and the inertial coordinate −ṙ(t)
frame G. The wind direction forms the angle ϕW with respect to ex and
defines the wind window (dotted curve). Note the arrows on the figure-eight A dynamic model of the described system can be derived
path showing an up-loop pattern, i.e., the wing is flying upward on the side from first principles, where the wing is assumed to be a point
of the path and downward in the middle. with the given mass. The tether is assumed to be straight
with a nonzero diameter. The aerodynamic drag of the tether
and the tether mass are added to the wing’s drag and mass,
the corresponding direction indicated by the trailing subscript respectively. The effects of gravity and inertial forces are also
letter. The ex axis is assumed to be parallel to the ground, considered. The wing is assumed to be steered by a change of
contained in the longitudinal symmetry plane of the GU, and the roll angle ψ(t), which is manipulated by the control system
the ez axis is perpendicular to the ground pointing upward via the line length difference δ(t). By applying Newton’s law
and the e y axis is such that it forms a right-hand system. of motion to the wing in the reference system L, we obtain
The wing’s position vector G p(t) can be expressed in the  
inertial frame using spherical coordinates (ϕ(t), ϑ(t), r (t))  F · eN 2 
ϑ̈ = − sin (ϑ) cos (ϑ) ϕ̇ 2
− ϑ̇ ṙ 
as (see Fig. 2)  rm r 

⎛ ⎞  F · eE 2 
r (t) cos (ϕ(t)) cos (ϑ(t))  ϕ̈ = + 2 tan (ϑ)ϑ̇ ϕ̇ − ϕ̇ ṙ  , (4)
 r m cos (ϑ) r 
G p(t) = ⎝ r (t) sin (ϕ(t)) cos (ϑ(t)) ⎠. (1)  F · eD
 
r (t) sin (ϑ(t))  r̈ = − + r ϑ̇ 2 + r cos2 (ϑ)ϕ̇ 2 
m
Note that all the three variables (ϕ(t), ϑ(t), r (t)) can be where m is the mass of the wing. The force F(t) consists of
measured directly with good accuracy by devices installed on contributions from gravity Fg (t), aerodynamic force Fa (t), and
the ground such as line angle sensors and motor encoders. the force exerted by the lines Fc (t). Note that for simplicity
The motion of the tethered wing is restricted to the wind of notation, we dropped the time dependence of the involved
window, a surface with (time-varying) radius r (t) confined by variables in (2) and (4). The force Fc (t), called traction force,
the ground plane (ex , e y ) and by a vertical plane containing opposes all other forces along the tether direction and can
the origin of G and perpendicular to the wind direction, which be influenced by the motors in the GU to control the tether
forms an angle denoted by ϕ W with respect to ex . If r (t) reeling. As mentioned above, the aerodynamic force Fa (t)
is kept constant, the wind window corresponds to a quarter can be influenced by the line length difference δ(t) of the
of a sphere. Otherwise, depending on the reeling speed ṙ (t) two steering lines, which is, as a first approximation, directly
of the tether, the wind window contains a larger or smaller related to the wing’s roll angle ψ(t). In particular, a change
surface area than a quarter sphere. For example, with ṙ (t) < 0, of δ(t) induces a change of the roll angle, which in turn
i.e., reeling in the tether, the wing is able to surpass the GU determines a change of orientation of the force vector Fa (t)
against the wind direction, thanks to the additional apparent acting on the wing. The full details on the derivation of this
wind speed induced by the reeling. model are available in [6].
In addition, we define a noninertial coordinate system Equation (4) gives an analytical expression for
.
L = (e N , e E , e D ), centered at the wing’s position (shown the point-mass model of the wing with six states,
in Fig. 2 too). The e N axis, or local north, is tangent to (ϕ(t), ϑ(t), r (t), ϕ̇(t), ϑ̇(t), ṙ (t)), two manipulated inputs
the sphere of radius r (t), on which the wing’s trajectory (δ(t), |Fc (t)|), and three exogenous inputs given by the
evolves, and points toward its zenith. The e D axis, called components of the wind vector W(t). Such a model has
local down, points to the center of the sphere (i.e., the GU), been widely used in the literature for the control design of
and hence, it is perpendicular to the tangent plane of the AWE systems (see [6], [16], [19], [20]).
sphere at the wing’s position. The e E axis, named local east, In a recent contribution [25] concerned with the autonomous
forms a right-hand system and spans the tangent plane together flight along figure-eight paths during the traction phase,

Authorized licensed use limited to: Consortium - Algeria (CERIST). Downloaded on June 08,2023 at 05:20:22 UTC from IEEE Xplore. Restrictions apply.
ZGRAGGEN et al.: AUTOMATIC RETRACTION AND FULL-CYCLE OPERATION FOR A CLASS OF AWE GENERATORS 597

the notion of the velocity angle γ has been introduced assuming crosswind flight conditions as performed during the

traction phase.
. v P (t) · e E (t)
γ (t) = arctan (5) During retraction, the tethers have to be recoiled onto the
v (t) · e N (t)
P
drums under minimal force such that only a small fraction
cos (ϑ(t))ϕ̇(t) of the previously generated energy is used. To achieve this
= arctan . (6)
ϑ̇(t) goal, the common strategy adopted with flexible wings, like
Thus, γ(t) is the angle between the local north e N (t) and the power kites, is to move the wing at the border of the wind
projection of the wing’s velocity vector v P (t) onto the tangent window, in a static angular position with respect to the GU,
plane spanned by the local north and east vectors, see (3). i.e., with constant or slowly varying ϕ and ϑ angles. This
In (6), the four-quadrant version of the arc tangent function represents quite a different flight condition with respect to
shall be used, such that γ (t) ∈ [−π, π]. the one assumed in (7). However, as discussed in the next
The velocity angle describes the flight conditions of the section, it can be shown that the model (7) can also, with
wing with just one scalar: as an example, if γ = 0, the wing some modifications, be used to describe the wing’s steering
is moving upward toward the zenith of the wind window, and if dynamics during the retraction phase, employing a slightly
γ = π/2, the wing is moving parallel to the ground toward the modified definition of the velocity angle (6) called regularized
local east. In addition, a control-oriented model for tethered velocity angle.
wings, originally proposed in [26] and refined in [25], has
been used for the control design of the traction phase III. AUTOMATIC R ETRACTION OF G ROUND -BASED
A IRBORNE W IND E NERGY S YSTEMS
γ̇ (t)  K (t)δ(t) + T (t) (7)
The problem of automatically retracting the wing during the
where reel-in phase involves two main tasks: reeling the tether on the
2 drum by controlling the line force and stabilizing the angular
ρC L (t)A|Wa (t)| 1 position of the wing at the border of the wind window. One of
K (t) = 1+ 2 (8a)
2mds E eq (t) our contributions is to show that the dynamical behavior of the
g cos (ϑ(t)) sin (γ (t)) wing’s elevation angle during retraction is almost linear so that
T (t) = + sin (ϑ(t))ϕ̇(t). (8b) standard linear control techniques can be applied. In addition,
|Wa (t)|
the reeling control can be considered as a decoupled problem
Equation (7) represents a simpler model, capturing only the that influences the position control system as a disturbance.
steering behavior of the wing, than the one represented in (4). We will present two different control strategies for the
In particular, note that the velocity angle (6) is a known non- problem of stabilizing the wing’s position during the retraction
linear function of the states in the point-mass model (4): from phase. The first approach, presented in Section III-A, exploits
the point of view of the control design, it can be considered as a regularized version of the velocity angle employed by the
an output of the system that can be used as feedback variable. traction phase controller. The resulting controller for the reel-
Equation (4) can be used to derive the model (7) through some in phase needs only minor changes with respect to the one
simplifying assumptions and manipulations, as shown in [25] employed in the traction phase. However, the regularized
and omitted here for brevity. velocity angle is computed on the basis of an estimate of the
In (7) and (8), the steering input, i.e., the line length wind direction and speed at the wing’s location. The second
difference of the steering lines, is denoted by δ(t), ρ is the approach, explained in Section III-B, is based on a simplified
air density, C L (t) is the aerodynamic lift coefficient, A is the model, derived in this paper, of the elevation dynamics of the
reference area of the wing, ds is the span of the wing, E eq (t) is
. tethered wing during the retraction. This approach has the
the equivalent efficiency of the wing, defined as E eq (t) = advantage of employing only directly measurable quantities
C L (t)/C D,eq (t), where C D,eq (t) represents the drag coefficient (the elevation angle ϑ and its time derivative), and hence,
of the wing and lines together, and g is the gravitational it does not need an estimate of the wind direction nor of the
acceleration. The apparent wind Wa (t) is defined as velocity angle. In Section III-C, we highlight the connections
Wa (t) = W(t) − v p (t) (9) between the two approaches and in Section III-D, we discuss
the reeling strategy.
where the incoming wind W(t) in the L frame can be
written as
⎛ ⎞ A. Retraction Control Based on the Regularized
−W0 cos (ϕ(t) − ϕ W ) sin (ϑ(t)) Velocity Angle
L W(t) = ⎝ −W0 sin (ϕ(t) − ϕ W ) ⎠ (10)
One of the main differences between the traction and
−W0 cos (ϕ(t) − ϕ W ) cos (ϑ(t))
retraction phases lies in the magnitude of the wing’s velocity
p
with W0 being the nominal wind speed (which can even- perpendicular to the line direction, denoted by v P (t). During
tually also be position dependent if a wind shear model is the retraction phase, v P (t) is low and mainly consists of the
p
considered [28]). reel-in speed ṙ (t). Thus, v P (t) is close to zero and the apparent
The model (7) has been validated through experimental wind speed is determined only by the wind speed W(t) and the
data at constant line length with good correspondence in a reel-in speed ṙ (t). In these conditions, the velocity angle γ as
wide range of operating conditions (see [25]). It was derived computed in (6) becomes undefined so that this variable is not

Authorized licensed use limited to: Consortium - Algeria (CERIST). Downloaded on June 08,2023 at 05:20:22 UTC from IEEE Xplore. Restrictions apply.
598 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 24, NO. 2, MARCH 2016

representative of the wing’s orientation anymore and cannot be


used for feedback control.
Recall that we assume for simplicity that the wind flow is
parallel to the ground, i.e., the (ex , e y ) plane and its direction
forms an angle ϕ W with respect to ex (see Fig. 2). It is also
assumed that the wing is designed so that it orientates itself
into the apparent wind, which means that the wing’s longi-
tudinal symmetry axis is aligned with the vector Wa (t) (9),
i.e., the wind direction during retraction, projected onto the
Fig. 3. Experimental data. Time courses of γ (t) (dashed line) and
tangent plane to the wind window at the wing’s location. This γ r (t) (solid line) during a transition from flying figure-eight paths in cross-
effect can be achieved by a wing equipped with a rudder or a wind conditions (up to approximately 282 s) to a position at the border of the
curved shape, like C-shaped surf kites. Thus, during retraction, wind window.
the component of Wa (t) in the tangent plane to the wind
window can be assumed to be equal to the wind velocity
projected on the same plane.
Under this assumption, we can define the orientation β(t)
of the wing using (10) as

. − L W(t) · e E (t)
β(t) = arctan (11)
− L W(t) · e N (t)

Fig. 4. Overview of the control approach that exploits the regularized velocity
sin (ϕ − ϕ W ) angle as a feedback variable.
= arctan (12)
sin ϑ cos (ϕ − ϕ W )
which is the angle between the local north e N and the With the regularized velocity angle (13), we can
longitudinal symmetry axis of the wing. now adopt a control approach for the retraction phase
From (12), one can see that β converges to ±π/2 similar to the one used for the traction phase described
if the wing approaches the border of the wind window, in [25].
i.e., when ϕ ≈ ϕ W ± π/2. An estimate of the wind direc- In particular, we consider a hierarchical control scheme
tion ϕ W , needed to compute the angle β, can be either obtained consisting of three nested loops, shown in Fig. 4. Note that
by measurements provided by ground based sensors or by the regularized velocity angle cannot be directly measured and
processing the measurements of the line force collected during needs to be estimated. For this purpose, line angle sensors
the traction phase (see [29]). installed on the GU can be used to measure the angular
The considerations presented so far lead to the idea of position of the wing. Using a Kalman filter with a linear free
extending the definition of the velocity angle γ by a reg- particle model, the velocity vector of the wing can then be
ularization term such that it can well represent the wing’s estimated and expressed in the L frame. Using (5), the velocity
orientation also for static positions at the border of the wind angle can finally be estimated (for details, see [30]), where also
window. In particular, we define the regularized velocity angle the additional use of on-board sensors is considered.
as, compared with (6) and (12) Besides the use of the regularized velocity angle as a

feedback variable, the main difference between the retraction
cos (ϑ)ϕ̇ + c sin (ϕ − ϕ W )
γ r = arctan (13) and the traction phases lies in the computation of the velocity
ϑ̇ + c sin ϑ cos (ϕ − ϕ W ) angle reference γrefr . Therefore, we will briefly recall the

where c > 0 is a scalar chosen by the control designer. equations describing the inner control loops only for the sake
In principle, the value of c should reflect the magnitude of of completeness (see [25] for details) and focus here on the
the absolute wind speed at the wing’s position divided by outer control loop, responsible for providing the velocity angle
the tether length, i.e., W0 /r , which might be not trivial to controller with a suitable reference.
obtain if no onboard wind speed sensors or ground wind Neglecting higher order effects and external disturbances,
profilers like LIDARs are present. However, in simulations the closed-loop actuation system can be modeled as a second-
and experiments, the system behavior resulted to be not order system
sensitive to this quantity, due to the relatively large line length δ̈m = ωcl
2
δref − 2ζcl ωcl δ̇m − ωcl
2
δm (14)
values (50–200 m) compared with the absolute wind
speed (3–6 m/s). where δm is the actuator’s position, δref is the actuator’s
Thus, according to (13), during the traction phase when position reference, and ωcl and ζcl are the natural frequency
the speed of the wing is significantly larger than the wind and damping, respectively, of the actuation control loop. The
speed W0 , we have γr ≈ γ, but during the retraction phase, steering deviation is then obtained as
when the wing speed approaches zero, γr still provides a
δ = K δ δm (15)
reasonable value, whereas the original velocity angle γ (6)
becomes undefined. A comparison between γ (t) and γr (t) where K δ is a known constant that depends on the mechanical
during a flight test is shown in Fig. 3. setup of the system, which defines the ratio between the

Authorized licensed use limited to: Consortium - Algeria (CERIST). Downloaded on June 08,2023 at 05:20:22 UTC from IEEE Xplore. Restrictions apply.
ZGRAGGEN et al.: AUTOMATIC RETRACTION AND FULL-CYCLE OPERATION FOR A CLASS OF AWE GENERATORS 599

actuator’s position and the effective difference of length of the we can rewrite the system dynamics in terms of angle errors
steering lines. In the case of the Swiss Kite Power prototype,
for example, K δ = 1. The velocity angle control loop consists γ r = γref
r
− γr (18)
of a proportional controller given by ϑ = ϑref − ϑ (19)
r  and of the position and velocity of the actuation system,
δref = K c γref − γr (16)
δm and δ̇m . To formulate the error dynamics, we need an
where the gain K c is chosen by the designer. intermediate step to include the dynamics of the angle ϑ.
As already mentioned, the goal of the retraction controller To this end, consider the apparent wind velocity vector pro-
is to stabilize the wing at a static position in terms of ϕ and ϑ jected onto the plane spanned by (e N , e E ). We denote such
at the border of the wind window, e.g., ϕ − ϕ W = ±π/2, and a projection, which can be computed by taking the first
p
at a given elevation angle ϑref . As seen in the previous section, two components of (9), as Wa . At the border of the wind
p
from (13), we have γ r = π/2 for a static position of the wing window, i.e., when ϕ − ϕ W = ±π/2, the component of Wa
with ϕ − ϕ W = π/2. This corresponds to a wing position on in the local north direction e N is given by r ϑ̇ only, since
the left of the wind window as seen from the GU. Similarly, the absolute wind results to be perpendicular to the local
if a position on the right of the wind window is considered, north. The quantity r ϑ̇ is, by definition of γ r (13), also
p
i.e., ϕ − ϕ W = −π/2, the regularized velocity angle becomes equal to |Wa | cos γ r, which can equivalently be written as
p
γ r = −π/2. For simplicity, we will now consider only |Wa | sin (π/2 − γ r ). Moreover, since the wing tends to align
positions on the left of the wind window for the retraction itself with the wind direction, π/2 − γ r is small, so that we
phase, i.e., ϕ − ϕ W = π/2 (the application to positions on the can linearize its trigonometric functions. Then, the dynamics
right of the wind direction is straightforward). of the elevation angle ϑ can be written as
Using the point-mass model (4) of the tethered wing, it can p

|Wa | π
be shown that there exist equilibrium points at the border of ϑ̇ = − γr . (20)
the wind window, whose values are a function (for a given r 2
wing) of the steering input δ and of the absolute wind speed. We can now derive the system dynamics pertaining to the
These equilibrium points can be computed as usual by setting closed-loop system when the regularized velocity angle is used
all time derivatives of the model states to zero and solving (4) as a feedback variable. In particular, using (7) and (14)–(20),
for a given steering input. In addition, they can also be found and setting x = [ ϑ, γ r , δm , δ̇m ]T (where T stands for
by numerical simulations of the point-mass model employing the matrix transpose operation), the following state-space
a constant steering input. This suggests that these equilibrium equations are obtained:
points are open-loop stable and have a nonempty region of ⎡ p p ⎤
attraction, as it is also revealed by commonly used analysis |Wa | |Wa |
⎢Kϑ r − 0 0 ⎥
techniques (see [31]). ⎢ p r p ⎥
⎢ 2 |Wa | |Wa | ⎥
Inspired by the above considerations, we propose the fol- ẋ = ⎢ K ϑ −K ϑ −K K δ 0 ⎥ x + w.
lowing feedback control strategy to compute a reference value ⎢ r r ⎥
⎣ 0 0 0 1 ⎦
for the velocity angle:
0 K c ωcl
2 −ωcl
2 −2ζcl ωcl
π   
γref
r
= K ϑ (ϑref − ϑ) + , K ϑ < 0 (17) Acl (21)
2
where ϑref is a reference elevation angle chosen by the user, For the sake of completeness, we briefly outline the
which should theoretically correspond to an equilibrium point derivation of (21). ϑ̇ is given by the time derivative of (19)
for the wing at the side of the wind window. From (17), and inserting (17) and (20). γ̇ r is given by the time derivative
one can note that if the elevation of the wing is smaller of (18) and inserting (7) and (17). By inserting the derivation
than the reference elevation, the velocity angle reference is of ϑ̇ described just above, and using the static relation-
smaller than π/2, thus demanding the wing to move toward the ship (15), an expression that is a function of x and T , see (8b),
zenith of the wind window. Contrarily, if the current elevation is obtained. The latter term is then embedded in the external
is larger than the reference one, we have γref r > π/2. This disturbance w. The derivation of δ̇m is straightforward, and
reference is saturated to γref ∈ [γmin , γmax ] to prevent the
r finally, δ̈m is given by (14) where δref is replaced with (16).
wing from turning away too much from the wind direction. In (21), the term K corresponds to the uncertain gain in (8a)
Such a situation could in fact give place to a transient in and depends on the system’s parameters as well as the wind
crosswind conditions, which would increase the traction force and the flight conditions. The term w ∈ R4 accounts for
unnecessarily. effects of gravity and apparent forces, i.e., the T in (8b),
The scalar gain K c for the velocity angle controller and as mentioned above, as well as for the forces exerted by the
the scalar gain K ϑ for the reference computation are chosen lines on the actuator. Equation (21) has time-varying uncertain
by the designer. Using (17) in the outer loop of the control linear dynamics characterized by the matrix Acl ( ), where
p
scheme (see Fig. 4), the resulting control system is linear = [K , |Wa |]. Upper and lower bounds for such parameters
(time varying) and the controller gains K ϑ and K c can be can easily be derived on the basis of the available knowledge
chosen such that robust stability is achieved in the presence of on the system. These bounds can be employed to compute
model uncertainty and different wind conditions. In particular, points i , i = 1, . . . , n v , such that ∈ conv( i ), where conv

Authorized licensed use limited to: Consortium - Algeria (CERIST). Downloaded on June 08,2023 at 05:20:22 UTC from IEEE Xplore. Restrictions apply.
600 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 24, NO. 2, MARCH 2016

denotes the convex hull. Then, the closed-loop system (21)


results to be robustly stable if there exists a positive definite
matrix P = P T ∈ R4x4 such that (see [32])
T
Acl ( i )P + P Acl ( i ) ≺ 0, i = 1, . . . , n v . (22)
Condition (22) can be checked using an LMI solver.
In Section IV, we show with simulations and experiments that
indeed a single pair (K c , K ϑ ) achieves robust stability of the Fig. 5. Overview of the control approach that exploits the elevation angle
as a feedback variable.
control system, as predicted by the described theoretical analy-
sis. The two scalar gains, i.e., the values of K c and K ϑ , can
be tuned at first using (7) and (17), and then via experiments. The model in (24) gives a direct relationship between the
As shown in Section IV, this approach is able to stabilize input δ and the elevation of the wing ϑ. We will now elaborate
the wing at the border of the wind window but is dependent a bit more on this result and its implications. As we can see
on an estimate of the wind speed and direction at the wing’s from (24), gravity and apparent forces have less influence
location. Since these might not be straightforward to obtain, with increasing tether length, since the linear acceleration
an alternative approach is presented in the next section, which remains constant such that the angular one is inversely pro-
relies only on directly measurable quantities. portional to the radius. The term ρ AC L /(2r mds )(1 + 1/E eq2 )

in (25) remains roughly constant during the retraction and


B. Retraction Control Based on Elevation Dynamics is specific to the employed wing. Similarly to what was
As an alternative to the regularized velocity angle, we done in the previous section for the approach based on
propose here to use the elevation angle ϑ and its rate ϑ̇ as the regularized velocity angle, the effects of variations in
feedback variables. The main advantage of such an approach the aerodynamic coefficients on the elevation control sys-
is a higher reliability, since the elevation is directly measured tem due to changing wind conditions can be evaluated by
(in our case, by means of a line angle sensor installed on the means of a robustness analysis, as detailed later on in this
GU) and there is no need to estimate the wind direction at the section.
wing’s location. The angular elevation rate ϑ̇ can be estimated Equation (24) also implies that a larger area-to-mass ratio
with a Kalman filter using a linear free particle model as A/m gives, in general, a higher gain C and that the steering
mentioned above (see [30]). We will carry out the controller’s gain of wings with similar aerodynamic coefficients but dif-
design on the basis of a new model that links the elevation ferent sizes should not change much, provided that they have
dynamics to the steering input, which we derive next. a similar A/m.
From (4), we can write the ϑ-dynamics as Finally, as mentioned earlier, there exist equilibrium points
at the border of the wind window such that the wing angular
F · eN 2
ϑ̈ = − sin (ϑ) cos (ϑ)ϕ̇ 2 − ϑ̇ ṙ . (23) position will converge at a constant elevation angle ϑ for a
rm r given constant input δ. This can be now seen by solving (24)
We consider the following assumptions. for δ, which results in an explicit link between the steady-state
Assumption 1 (Steady State): The wing is at a steady-state values of ϑ and δ
angular position at the border of the wind window. 
2
Assumption 2 (Small Roll Angle): The roll angle ψ of the 2mgds cos (ϑ) E eq 1
δ= . (26)
wing is sufficiently small, such that its trigonometric functions ρ AC L E eq + 1 |Wa |W0 sin (ϕ)
2

can be linearized. 
Assumption 1 implies that the sum of the forces acting on However, the link given by (26) cannot be used in an open-loop
the wing in the local east direction, e E , is zero and that the control approach due to the presence of model uncertainties
angular velocities of the wing are small. Thus, effects from and changing wind conditions which render a feedback con-
apparent forces are small. Moreover, this also implies that troller for the elevation angle necessary.
the wing’s longitudinal axis is aligned with the apparent wind Exploiting the model (24), such a controller can be
direction. Assumption 2 is also reasonable, since, for example, designed by considering again a hierarchical control system,
during our test flights, the roll angle was within ±18◦ . We can now consisting of only two nested loops, the low-
now state our result concerned with the wing’s model. level actuation control loop, and the elevation controller,
Proposition 1: Let Assumptions 1–2 hold. Then, the ϑ shown in Fig. 5.
dynamics (23) can be written as To design the controller, (24) is first linearized around an
equilibrium point, which serves as reference position ϑref .
g cos (ϑ) + 2ϑ̇ ṙ As pointed out in Section III-A, such an equilibrium point
ϑ̈ = −Cδ − (24)
r can be found using the point-mass model (4). The resulting
where linear system is given by
⎡ ⎤  
ρ AC L 1 0 1 0
C= 1 + 2 W0 sin (ϕ − ϕ W )|Wa |. (25) ẋ = ⎣ g sin (ϑref ) 2ṙ ⎦ x +
 
u (27)
2r mds E eq − −C
Proof: See the Appendix. r r

Authorized licensed use limited to: Consortium - Algeria (CERIST). Downloaded on June 08,2023 at 05:20:22 UTC from IEEE Xplore. Restrictions apply.
ZGRAGGEN et al.: AUTOMATIC RETRACTION AND FULL-CYCLE OPERATION FOR A CLASS OF AWE GENERATORS 601

where x = [ ϑ, ϑ̇]T and u = δ. The tracking error comes from how the force component in ϑ direction, F · e N , is
in ϑ and ϑ̇ are given as calculated. In (24), this component is computed by considering
the apparent wind in the tangent plane at the wing’s position,
ϑ = ϑref − ϑ (28)
i.e., W0 sin (ϕ − ϕ W ) + r cos (ϑ)ϕ̇, where W0 sin (ϕ − ϕ W ) is
ϑ̇ = ϑ̇ref − ϑ̇ (29) the dominating factor (see the Appendix). On the other hand,
p
where the reference values correspond to a static angular the corresponding term in (20) is Wa , which corresponds,
position, i.e., ϑ̇ref = 0. assuming a static angular position at the border of the wind
We use a state feedback controller K SF of the form window and constant line length, to Wa  W0 sin (ϕ − ϕ W ).
In summary, it has to be noted that the structure of the
z = −K SF x (30) two models is the same, which explains why the corresponding
where z = δref and K SF = [k1SF k2SF ] is a vector of feedback controllers have similar qualitative behavior, as it will be
gains that can be designed by means of standard techniques shown in Section IV, but with quite marked performance
like pole-placement or linear-quadratic (LQ) regulation. Again, differences in tracking the reference elevation ϑref .
it can be shown that there exists a matrix K SF for which the
system is robustly stabilized in the presence of the uncertain D. Reeling
time-varying parameters. A robustness analysis can be carried As mentioned above, the reeling can be considered, from
out similarly to the one in Section III-A; the corresponding the point of view of the position control system, as an
closed-loop dynamics are given, using (14), (15), (27)–(30), external disturbance since its main influence is on the mag-
and x = [ ϑ, ϑ̇, δm , δ̇m ], by nitude of the apparent wind speed and all other effects
⎡ ⎤ are comparably small. This is the reason why both traction
0 1 0 0
⎢ g sin (ϑref ) 2ṙ ⎥ and retraction controllers can be designed independently
⎢ − −C K δ 0 ⎥  from the reeling speed control. For simplicity, we there-

ẋ = ⎢⎢ r r ⎥ x + w.
⎥ fore adopt a simple reeling control scheme for both phases
⎣ 0 0 0 1 ⎦
of the power cycle, by setting a torque reference on the
−ωcl
2 k SF
1 −ωcl
2 k SF
2 −ωcl
2 −2ζcl ωcl
   (31) generators.
Acl During the traction phase, the torque reference is chosen
Here, the uncertain time-varying parameters are given with a feedback strategy that aims to achieve the optimal reel-
by = [r, ṙ, C]. out speed [5]. In particular, assuming a steady-state reeling,
i.e., constant speed, where the optimal traction force has to be
C. Discussion matched by the motor torque, we have
We presented two control approaches for the retraction
Tm = Fc∗rd (33)
phase, one based on a regularized version of the velocity
angle γ and one based on the ϑ-dynamics derived from the where Tm is the torque applied by the motor, rd is the radius
first-principle model (4). In the latter, we exploit a direct of the drum, and Fc∗ is the optimal traction force for maximum
link between the input δ and the angular acceleration ϑ̈, power production for a given wind situation. A simplified
while the first approach does not explicitly consider the ϑ model of the traction force Fc has been first introduced
dynamics and relies on the turning rate γ̇ instead. For the in [5] and then subsequently refined in several contributions
sake of comparison, also in the first approach, one can derive (see [33])
the angular acceleration ϑ̈, in particular, by taking the time
2
derivative of (20) and combining it with (7), and assuming Fc (t) = |Fc (t)| = C War (34)
that the apparent wind velocity Wa and the value of ϕ are
constant with

ρ AC L 1 g cos (ϑ) 3
ϑ̈ = − 1 + 2 |Wa |2 δ − . (32) 1 2 1 2
2r mds E eq r C = ρ AC L E eq 1 + 2 (35)
2 E eq
Comparing this equation with the one derived from the
model (24), one can see a few differences. First, the second where ρ is the air density, A is the wing reference area,
term in the right-hand side of (32) does not contain the term C L is the lift coefficient, E eq is the equivalent efficiency,
related to apparent forces. This comes from the fact that the and War is the apparent wind vector component in tether
ϑ-dynamics in (20) do not consider the influence of the reeling direction, consisting of the contributions of the absolute wind
speed ṙ . The term related to gravity is the same since we speed W and of the reeling speed ṙ . More specifically, War can
assume γ r ≈ π/2 for the retraction. Note that as one would be derived by using the third entry of (9), given in turn
expect, for both models, the influence of the additive terms by (3) and (10), for which the wing’s angular position, the
on the angular acceleration becomes smaller for longer tether wind speed, and the reeling speed are needed. It can be shown
length. that the reeling speed that yields the maximum generated
The gain relating the input δ to ϑ̈, denoted by C in (24), is power during the reel-out phase (corresponding to the optimal
quite similar to the corresponding gain in (32). The difference traction force Fc∗ ) is equal to one third of the wind speed

Authorized licensed use limited to: Consortium - Algeria (CERIST). Downloaded on June 08,2023 at 05:20:22 UTC from IEEE Xplore. Restrictions apply.
602 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 24, NO. 2, MARCH 2016

TABLE I TABLE II
S YSTEM PARAMETERS C ONTROL PARAMETERS

in tether direction (see [5]). By indicating such an optimal


reel-out speed with r ∗ , we can then express (War )2 in (34) as
r 2
Wa = (W r − ṙ )2 (36)
= (3ṙ ∗ − ṙ )2 (37)
.
where Wr = |W| cos (ϕ − ϕ W ) cos (ϑ) is the wind speed in
tether direction. Thus, the motor torque required to achieve a
given reel-out speed ṙ is
Tm = C rd (3ṙ ∗ − ṙ )2 . (38)
Then, considering that ṙ can be measured quite accurately
using the rotary position sensors of the ground generators and
that a reasonably good estimate of C during the traction phase
can be easily obtained from the knowledge available on the
wing as well as from experimental data (see [25]), we can set
the motor torque to Fig. 6. Simulation results. Typical 3-D trajectory (black) and its projec-
tion (gray) on the ground of the tethered wing during one flown power cycle.
Tm = 4C rd ṙ 2 (39)
hence obtaining ṙ = ṙ ∗
as a unique positive solution of (38).
approach of Section III-B, an LQ regulator with weighting
Furthermore, with a simple analysis of the reeling dynamics
matrices equal to the identity matrix was used.
(involving the generator’s inertia and viscous friction), it can
With the employed control gains, the robustness
be shown that such a solution is an asymptotically stable
analyses described in Sections III-A and III-B indicated
steady state when the feedback reeling strategy (39) is used.
that the closed-loop system is stable for wind speeds in the
In addition, we included a lower and an upper bound on the
range of [1, 15] m/s, a lift coefficient in the range of [0.4, 1],
torque reference to avoid wing stall and mechanical overload
and an aerodynamic efficiency in the range [2, 8], covering
of the system, respectively.
a quite broad variety of operating conditions. Indeed, we
During the retraction, a constant torque reference is chosen
employed the gains indicated in Table II in all our simulations
to achieve a high reel-in speed, to increase the duty-cycle of
and experimental tests, in the presence of varying and gusty
the overall power generation scheme.
wind conditions, and also with different kite sizes, with
Indeed, the interplay between the wing dynamics and reel-
satisfactory performance.
ing speed could be exploited using a multivariable control
In Fig. 6, a typical trajectory of the wing from launch
technique with the aim to optimize the power output. If an
until the end of the first power cycle is shown. At first, the
additional actuator to change the pitch angle of the wing is
wing is flown in crosswind conditions, flying figure-eight paths
also present, i.e., allowing one to change the lift and drag
until it reaches the maximum tether length of 150 m, using
coefficients of the wing, the efficiency of the system could
the controller described in [25]. Then, the retraction phase
be further increased. These topics are not considered in this
is started using either the controller based on the regularized
paper, but they represent further research directions.
velocity angle (16), (17) or the feedback controller (30), while
the tether is reeled in until a length of 50 m is reached.
IV. R ESULTS At that point, the traction phase controller of [25] is used
We first compare the proposed control approaches for the again to complete the power cycle. In Fig. 7, the time courses
retraction phase in simulation, employing the nonlinear point- of the position angles ϕ and ϑ during one power cycle
mass model for tethered wings (4) and considering ϕ W = 0. for both control approaches are shown. Around 73 s, the
The main system and controller parameters are shown controller switches from traction to retraction and tracks the
in Tables I and II, respectively. The terms relating to γ r apply reference ϑref = 1 rad. Note that ϕ becomes slightly larger
only to the approach from Section III-A; for the state feedback than π/2 due to the reel-in speed, indicating that the wing

Authorized licensed use limited to: Consortium - Algeria (CERIST). Downloaded on June 08,2023 at 05:20:22 UTC from IEEE Xplore. Restrictions apply.
ZGRAGGEN et al.: AUTOMATIC RETRACTION AND FULL-CYCLE OPERATION FOR A CLASS OF AWE GENERATORS 603

Fig. 9. Experimental results. Wing path projected onto the ground,


corresponding to Fig. 10. The wind direction was roughly ϕW ≈ −0.4 rad.

elevation angle and its rate, which are both measured with
good accuracy, as feedback variables. On the other end,
the first approach that we presented exploits the regularized
velocity angle, whose estimate can be inaccurate due to the
Fig. 7. Simulation results. Time courses of ϕ (dashed line) and ϑ (solid line) uncertainty in the wind speed estimation, i.e., the tuning
of one power cycle with a reel-in speed of 2.5 m/s and W0 = 5 m/s. parameter c in (13), instead of the elevation rate, with the
(a) Using the controller based on the regularized velocity angle. (b) Using
the controller based on the elevation dynamics.
consequent performance degradation. This is shown in detail
in Fig. 8, where the average tracking error of one retraction
phase for different reel-in speeds and different wind speeds,
respectively, are plotted for the two approaches.
We also carried out real-world experiments using the
Swiss Kite Power prototype shown in Fig. 1. The wing
employed in the tests presented here was a three-line
A IRUSH O NE surf kite with an area of 9 m2 . We collected
data for about 2 h of autonomous operation with the controller
based on the regularized velocity angle and about 1 h with the
controller based on the elevation dynamics.
In Figs. 9 and 10, the results of experimental test flights with
the retraction control strategy proposed in Section III-A, which
employs the regularized velocity angle as feedback variable,
are shown. In Fig. 10(a), the wing path during a power cycle
in the (ϕ, ϑ)-plane is shown. The wing is controlled to fly
along the figure-eight trajectories until it reaches the maximum
tether length and then flies horizontally to the border of the
wind window. Such a transient phase can be achieved by
setting a new target point for the traction controller at the
border of the wind window. Then, the retraction controller
Fig. 8. Average ϑ tracking error for the control system based on the stabilizes the wing during the reel-in of the tether. Once at
regularized velocity angle (solid line) and the one based on the elevation
dynamics (dashed line) during one retraction phase. (a) For different reel-in
the minimum tether length, the wing turns back to fly figure-
speeds with W0 = 5 m/s. (b) For different wind speeds W0 with a reel-in eight crosswind paths. In Fig. 10(b), the velocity angle and
speed of 2.5 m/s. its reference are shown, and in Fig. 10(c), the corresponding
time courses of ϕ and ϑ are shown. Note that the wing flies
downward to a low ϑ angle when starting a new traction
surpasses the GU location against the wind. Around 138 s, phase. This is due to the increasing wing speed and rather
the controller switches from retraction to traction and the small K c gain used for this maneuver. This problem can be
wing starts again flying figure-eight paths in crosswind alleviated by increasing the steering gain for this phase, as
conditions. we show later in Fig. 11(a). A projection of the wing path
Both control approaches lead to qualitatively similar results, on the ground plane can be observed in Fig. 9. Note that the
as it can be observed from Fig. 7. The main noticeable wing surpasses the GU upwind, since it reaches a negative
difference is the tracking of the ϑ reference during retrac- position in the ex direction. The average wind speed was
tion, which is better achieved by the approach that exploits approximately 4.6 m/s. The time course of the wind measured
the newly derived model (24) for the elevation dynamics. roughly 5 m above the ground can be seen in Fig. 10(d). The
This is expected, since this controller directly employs the resulting traction force on the main line during the power cycle

Authorized licensed use limited to: Consortium - Algeria (CERIST). Downloaded on June 08,2023 at 05:20:22 UTC from IEEE Xplore. Restrictions apply.
604 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 24, NO. 2, MARCH 2016

Fig. 10. Experimental results of one power cycle obtained using the retraction control strategy based on the regularized velocity angle. (a) Wing
trajectory in ϕ and ϑ. The wind direction was roughly ϕW ≈ −0.4 rad with an average wind speed of 4.6 m/s. (b) Time courses of γ (solid curve) and
γref (dashed curve). At roughly t ∈ [558 s, 567 s], the regularized version of γ (13) is used for feedback control. (c) Time courses of ϕ (solid curve) and
ϑ (dashed curve). (d) Wind speed (dotted curve) and a 1-min average wind speed (solid curve). (e) Time course of the traction force on the main line (solid curve).
(f) Time course of tether length r (solid curve).

is shown in Fig. 10(e). It can be seen that there is a significant line is visible in Fig. 11(e), where it can be noted that during
drop in traction force during the retraction phase, as expected the retraction phase, the force drops by a factor of about 2.5.
from the considerations above, leading to a positive net energy The resulting tether length during the power cycle is shown
output of the system. The time course of the tether length can in Fig. 11(f).
be seen in Fig. 10(f). A movie of the autonomous power cycles There are a few notable differences between
is available online [34]. Figs. 10 and 11. To decrease the traction force on the
Fig. 11 shows the results of experimental test flights, where lines and the ϕ position overshoot behind the GU against the
the approach based on the elevation dynamics has been used, wind, the pitch angle of the wing was slightly increased in
with the same A IRUSH O NE 9 m2 kite. the experiment shown in Fig. 11 compared with that in the
In Fig. 11(a), the wing path during a power cycle in experiment shown in Fig. 10, resulting in a lower efficiency
the (ϕ, ϑ)-plane is shown. Again, the retraction controller of the wing. In addition, the gain K c was kept at a higher
stabilizes the wing at the border of the wind window during value once the new traction phase starts until the wing is in
the reel-in of the tether. Once at the minimum tether length, a downwind position (see Table II). This compensates for the
the wing turns back to fly toward a downwind position decrease in the wing’s steering gain, see (8a), and prevents
using the traction controller [25]. The time course of the the wing from flying to a low elevation once the traction
wing elevation is shown in Fig. 11(b) together with its phase starts, comparing Figs. 10(a) and 11(a). In addition,
reference when the retraction controller was active (roughly as expected, the controller based on the elevation dynamics
at [510 s, 518 s]), clearly showing the closed-loop track- shows a slightly better tracking performance for ϑ. This can
ing behavior. In Fig. 11(c), the corresponding time courses be seen by comparing Fig. 10(c) (between 560 and 570 s)
of ϕ and ϑ during the power cycle are shown. One can and Fig. 11(c) (between 510 and 520 s).
see that the elevation-based retraction controller corrects the In Fig. 12(a) and (b), a comparison of the traction force
low ϑ position of the wing (starting roughly at 510 s) toward between the actual measurements during one power cycle
ϑref = 1 rad. The wind speed was approximately 5 m/s, and the simplified traction force model (34) is shown.
see Fig. 11(d). The corresponding traction force on the main In Fig. 13 (a) and (b), a comparison of the mechanical power

Authorized licensed use limited to: Consortium - Algeria (CERIST). Downloaded on June 08,2023 at 05:20:22 UTC from IEEE Xplore. Restrictions apply.
ZGRAGGEN et al.: AUTOMATIC RETRACTION AND FULL-CYCLE OPERATION FOR A CLASS OF AWE GENERATORS 605

Fig. 11. Experimental results of one power cycle obtained using the retraction control strategy based on the elevation dynamics. (a) Wing trajectory
in ϕ and ϑ. The wind direction was roughly ϕW ≈ 0.5 rad with an average wind speed of 5 m/s. (b) Time course of elevation ϑ (solid curve) and its reference
during the retraction phase (dashed curve). (c) Time courses of ϕ (solid curve) and ϑ (dashed curve). (d) Wind speed (dotted curve) and the 1 min average
wind speed (solid curve). (e) Time course of the traction force on the main line (solid curve). (f) Time course of tether length r (solid curve).

on the main line is shown, using the same model. This model situation where there is a relatively strong wind at ground level
has been widely used to estimate and optimize the power and much weaker wind at the wing’s position.
output of AWE systems during the traction phase, as well Despite the mentioned approximations, the two plots
as to support economical considerations for AWE generators. in Fig. 12 generally show a good correspondence during the
To carry out such a comparison, the lift coefficient and equiv- traction phase with the tendency of the simplified equation (34)
alent efficiency were estimated using a fraction of the data set. to slightly underestimate the traction force. During the retrac-
These values can change even for the same wing if different tion phase, the assumptions made in [33] do not hold anymore
bridling setups are used. In Fig. 12(a), the values are C L = 0.8 and the model exhibits a larger deviation, see Fig. 12(a).
and E eq = 3.7, whereas in Fig. 12(b), they are C L = 0.8 In addition, it has to be noted that our setup was not optimized
and E eq = 3.2. In addition, to estimate the magnitude for power production and thus the drop in traction force during
of the absolute wind projected along the lines’ direction, the retraction is not as large as the one that could probably
i.e., War in (34), one needs (as mentioned in Section III-D) the be achieved using, for example, ad hoc designed kites with
wing’s angular position, the absolute wind speed vector at the stronger depower capabilities and eventually additional control
wing’s position, and the reeling speed. While the kite’s angular strategies to adjust the kite’s pitch during reel-in. In Fig. 12(b),
position with respect to the ground and the reeling speed a drop in wind speed and a lower reel-in speed led to a good
were measured quite accurately with the available sensors, the matching of the traction force during the retraction phase.
absolute wind speed vector at the wing’s position was not To this regard, it also has to be noted that the magnitude of
available: we thus employed an estimate by assuming that the the traction force during the retraction depends on the reel-in
wind velocity measured roughly 5 m above the ground corre- speed: a larger reel-in speed results in a higher apparent wind
sponded to the one at the wing’s location. This approximation speed, which has a squared dependence on the traction force
should lead in general to an underestimate of the wind speed at (assuming that all the other quantities remain the same). The
the wings’s location and thus an underestimate of the traction spike in the line force in Fig. 12(b) given by the simplified
force, since typically an increasing wind speed profile above model (34) at roughly 475 s could not exactly be explained
the earth surface is experienced. However, transients and gusts by the available data; it is most probably caused by a wind
might also give rise, for short periods of time, to the opposite drop at the wing’s location, which is not seen by the ground

Authorized licensed use limited to: Consortium - Algeria (CERIST). Downloaded on June 08,2023 at 05:20:22 UTC from IEEE Xplore. Restrictions apply.
606 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 24, NO. 2, MARCH 2016

V. C ONCLUSION
We proposed two different approaches to design a feedback
controller for the retraction phase of an AWE system with
ground-based generation, where the tether is recoiled onto
the drums. Together with a previously proposed traction con-
troller and with a torque-based reeling control strategy, the
approaches presented here have been used to achieve fully
autonomous power cycles.
The two approaches were compared in simulation employ-
ing a nonlinear point-mass model and in real experiments
using the Swiss Kite Power prototype. Both approaches were
able to stabilize the wing in a position at the border of the
wind window and can be used to fly complete power cycles
with a tethered wing. The approach based on the elevation
dynamics is more promising since it relies only on directly
measured variables.
For both approaches, only few parameters that can be
intuitively tuned are involved in the design. The approaches
Fig. 12. Time courses of the mechanical power on the main line (solid curve)
and the corresponding mechanical power using the traction force model (34) employ the steering deviation as a control input and can
(dashed-dotted curve). (a) Corresponding to Fig. 10. (b) Corresponding stabilize the wing’s elevation robustly against different tether
to Fig. 11. lengths and reeling speeds. Hence, the latter can still be
optimized to maximize the energy output of the system.
The presented automatic controllers for the retraction phase
are two among the few, which have so far been proven to
work on real prototypes. Future research on this topic can be
devoted to the power cycle optimization using multivariable
approaches and to the inclusion of active pitch strategies, as
considered in [27].

A PPENDIX
P ROOF OF P ROPOSITION 1
For the sake of simplicity of notation, we drop the depen-
dence of time-varying variables and parameters on t.
The components of the force F in (4) in ϑ direction are
given by the gravitational force Fg and the aerodynamic
force Fa . The gravitational force can be expressed in the local
frame L as
⎛ ⎞
−mg cos (ϑ)
Fig. 13. Time courses of the traction force on the main line (solid curve) L Fg = ⎝ 0 ⎠ (40)
and the traction force model (34) (dashed-dotted curve). (a) Corresponding mg sin (ϑ)
to Fig. 10. (b) Corresponding to Fig. 11.
with m being the mass of the wing plus the added mass of the
based anemometer, that led to a short reel-in maneuver during tether and g is the gravitational acceleration. The aerodynamic
the traction phase to keep a minimum tension on the lines, force is given as
comparing Fig. 11(d) and (f) at about 475 s. More specifically,
the short reel-in transient increases War in the simplified model, Fa = FL e L + FD,eq eW (41)
while still assuming the same wind speed, and thus leads to
where FL is the lift force and FD,eq the equivalent drag force
an overestimate of the traction force with respect to the actual
also including the tether drag
measurement. Such events are quite ordinary in the presence
of relatively gusty wind like the one we encountered in most 1
of the experiment sessions. The time course of the mechanical FL = ρ AC L |Wa |2 (42)
2
power on the main line is compared with the simplified model 1
in Fig. 13(a) and (b). It can be noted that the average power FD,eq = ρ AC D,eq |Wa |2 . (43)
2
values are quite consistent during the traction phase and that
the simplified model is subject to lower variability since it does In (42) and (43), ρ is the air density, A is the effective area
not consider the changing wing speeds during the figure-eight of the wing, C L and C D,eq are the lift coefficient and equiva-
pattern. lent drag coefficient, and Wa is the apparent wind velocity.

Authorized licensed use limited to: Consortium - Algeria (CERIST). Downloaded on June 08,2023 at 05:20:22 UTC from IEEE Xplore. Restrictions apply.
ZGRAGGEN et al.: AUTOMATIC RETRACTION AND FULL-CYCLE OPERATION FOR A CLASS OF AWE GENERATORS 607

The vectors e L and eW in (41) can be expressed in the Using (50) together with (42)–(48), we obtain
L frame as
⎛ ⎞ ρ AC L 1
cos ξ − sin ξ 0 F · eN = 1 + 2 (W0 sin (ϕ − ϕ W )
2ds E eq
L e L (t) = ⎝ sin ξ cos ξ 0⎠
0 0 1 + r cos (ϑ)ϕ̇)|Wa |δ −mg cos (ϑ). (53)
⎛ ⎞
cos ψ cos η sin α Equation (23), by Assumption 1, can now be rewritten as
· ⎝cos ψ sin η sin α + sin ψ cos α ⎠ (44)
− cos ψ cos η cos α g cos (ϑ) + 2ϑ̇ ṙ
⎛ ⎞⎛ ⎞ ϑ̈ = −Cδ − (54)
r
cos ξ − sin ξ 0 − cos α
L e W (t) = ⎝ sin ξ 0⎠ ⎝ ⎠
cos ξ 0 (45) where

0 0 1 − sin α ρ AC L 1
C = 1+ 2 W0 sin (ϕ − ϕ W )|Wa |. (55)
where α is the angle between the apparent wind and the 2r mds E eq
tangent plane (e N , e E ) and ψ is the roll angle of the wing,

which is a function of the steering input δ
R EFERENCES

δ [1] M. Diehl, R. Schmehl, and U. Ahrens, Eds., Airborne Wind Energy


ψ = arcsin (46)
ds (Green Energy and Technology). Berlin, Germany: Springer-Verlag,
2014.
η is given by (see [35]) [2] L. Fagiano and M. Milanese, “Airborne wind energy: An overview,” in
Proc. Amer. Control Conf., Montreal, QC, Canada, 2012, pp. 3132–3143.
[3] Makani Power Inc., Alameda, CA, USA. (Sep. 2013). [Online].
η = arcsin (tan ( α) tan (ψ)) (47) Available: http://www.makanipower.com/
[4] Ampyx Power, The Hague, The Netherlands. (Sep. 2013). [Online].
and ξ is the heading of the wing, which is given by the Available: http://www.ampyxpower.com/
apparent wind vector Wa , defined in (9), and can be written as [5] M. L. Loyd, “Crosswind kite power,” J. Energy, vol. 4, no. 3,

pp. 106–111, May 1980.
−Wa · e E [6] M. Canale, L. Fagiano, and M. Milanese, “High altitude wind energy
ξ = arctan (48) generation using controlled power kites,” IEEE Trans. Control Syst.
−Wa · e N

Technol., vol. 18, no. 2, pp. 279–293, Mar. 2010.
W0 sin (ϕ − ϕ W ) + r cos (ϑ)ϕ̇ [7] Swiss Kite Power, Windisch, Switzerland. (Sep. 2013). [Online].
= arctan . (49) Available: http://www.swisskitepower.ch/
W0 sin (ϑ) cos (ϕ − ϕ W ) + r ϑ̇ [8] Twingtec AG, Dübendorf, Switzerland. (Jun. 2014). [Online]. Available:
http://twingtec.ch
The assumption underlying (48) is that the wing’s longitudinal [9] EnerKite GmbH, Berlin, Germany. (Sep. 2013). [Online]. Available:
symmetry axis is always contained in the plane spanned http://www.enerkite.de/
by the vectors Wa and p and is common in the field of [10] Kitenergy S.r.l., Turin, Italy. (Sep. 2013). [Online]. Available:
http://www.kitenergy.net/
AWE [19], [22], [36]. [11] SkySails GmbH & Co., Hamburg, Germany. (Sep. 2013). [Online].
Thus, the force F in e N direction can be computed as Available: http://www.skysails.info/power/
[12] Windlift, Inc., Raleigh, NC, USA. (Sep. 2013). [Online]. Available:
F · e N = FL (cos (η) sin ( α) cos (ξ ) http://www.windlift.com/
[13] e-Kite, Barneveld, The Netherlands. (Jun. 2014). [Online]. Available:
− sin (η) sin ( α) + cos ( α)ψ) sin (ξ )) http://www.e-kite.com
[14] I. Argatov and R. Silvennoinen, “Structural optimization of the pump-
− FD,eq cos ( α) cos (ξ ) − mg cos (ϑ). (50) ing kite wind generator,” Structural and Multidisciplinary Optimiza-
tion, vol. 40, nos. 1–6, pp. 585–595, 2010. [Online]. Available:
For more details and a formal definition of the components http://dx.doi.org/10.1007/s00158-009-0391-3
of F, see [25]. [15] C. Jehle and R. Schmehl, “Applied tracking control for kite power
systems,” J. Guid., Control, Dyn., vol. 37, no. 4, pp. 1211–1222, 2014.
By Assumption 1 and considering the equilibrium of the [16] M. Canale, L. Fagiano, and M. Milanese, “Power kites for wind
lift and drag force in the direction of the wing’s heading ξ , energy generation,” IEEE Control Syst. Mag., vol. 27, no. 6,
projected on the tangent plane to the wind window at the pp. 25–38, Dec. 2007.
[17] B. Houska and M. Diehl, “Robustness and stability optimization of
wing’s location, we have (see [33]) power generating kite systems in a periodic pumping mode,” in Proc.
IEEE Int. Conf. Control Appl. (CCA), Sep. 2010, pp. 2172–2177.
sin ( α) C D,eq . 1
= = (51) [18] E. J. Terink, J. Breukels, R. Schmehl, and W. J. Ockels, “Flight dynamics
cos ( α) CL E eq and stability of a tethered inflatable kiteplane,” J. Aircraft, vol. 48, no. 2,
pp. 503–513, Mar./Apr. 2011.
where E eq is the equivalent efficiency of the wing. By (51), [19] A. Ilzhöfer, B. Houska, and M. Diehl, “Nonlinear MPC of kites
under varying wind conditions for a new class of large-scale wind
we can see that α is small for a reasonable wing efficiency power generators,” Int. J. Robust Nonlinear Control, vol. 17, no. 17,
of 4–6. pp. 1590–1599, Nov. 2007.
By Assumption 2, (46), and (51), we see that (47) [20] J. H. Baayen and W. J. Ockels, “Tracking control with adaption of kites,”
simplifies to IET Control Theory Appl., vol. 6, no. 2, pp. 182–191, 2012.
[21] P. Williams, B. Lansdorp, and W. J. Ockels, “Optimal crosswind towing
1 1 and power generation with tethered kites,” J. Guid., Control Dyn.,
η= ψ= δ (52) vol. 31, no. 1, pp. 81–93, Jan. 2008.
E eq E eq ds [22] B. Houska and M. Diehl, “Optimal control for power generating
kites,” in Proc. Eur. Control Conf. (ECC), Kos, Greece, Jul. 2007,
where ds is the span of the wing. pp. 3560–3567.

Authorized licensed use limited to: Consortium - Algeria (CERIST). Downloaded on June 08,2023 at 05:20:22 UTC from IEEE Xplore. Restrictions apply.
608 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 24, NO. 2, MARCH 2016

[23] S. Costello, G. François, and D. Bonvin, “Real-time optimization for Lorenzo Fagiano (M’07) received the M.S. degree
kites,” in Proc. 5th IFAC Workshop Periodic Control Syst., Caen, France, in automotive engineering and the Ph.D. degree
Jul. 2013, pp. 64–69. in information and systems engineering from the
[24] M. Diehl, “Real-time optimization for large scale processes,” Politecnico di Torino, Turin, Italy, in 2004 and
Ph.D. dissertation, Dept. Natural Sci. Math., Heidelberg Univ., 2009, respectively.
Heidelberg, Germany, Jun. 2001. He was with the Fiat Research Centre, Turin,
[25] L. Fagiano, A. Zgraggen, M. Morari, and M. Khammash, “Automatic in 2005, where he was involved in active
crosswind flight of tethered wings for airborne wind energy: Modeling, vehicle systems. In 2007, he was a Visiting Scholar
control design, and experimental results,” IEEE Trans. Control Syst. with the Katholieke Universiteit Leuven, Leuven,
Technol., vol. 22, no. 4, pp. 1433–1447, Jul. 2014. Belgium. From 2009 to 2010, he was a
[26] M. Erhard and H. Strauch, “Control of towing kites for seago- Post-Doctoral Researcher with the Politecnico di
ing vessels,” IEEE Trans. Control Syst. Technol., vol. 21, no. 5, Torino. From 2010 to 2012, he was a Visiting Researcher with the University
pp. 1629–1640, Sep. 2013. of California at Santa Barbara, Santa Barbara, CA, USA, and a Senior
[27] B. Galletti, M. Buffoni, J. Ferreau, L. Fagiano, and M. Mercangoez, Researcher with the Automatic Control Laboratory, ETH Zurich, Zurich,
“Active pitch control of tethered wings for airborne wind energy,” in Switzerland, from 2012 to 2013. He is currently a Senior Scientist with
Proc. IEEE 53rd Annu. Conf. Decision Control (CDC), Los Angeles, ABB Switzerland Ltd., Corporate Research, Baden-Dättwil, Switzerland.
CA, USA, Dec. 2014, pp. 4893–4898. His current research interests include switchgear, airborne wind energy,
[28] M. L. Ray, A. L. Rogers, and J. G. McGowan, “Analysis of wind shear model predictive control, and experimental modeling.
models and trends in different terrains,” in Proc. AWEA Windpower Dr. Fagiano was a recipient of the IEEE T RANSACTIONS ON S YSTEMS
Conf., Pittsburgh, PA, USA, Jun. 2006. T ECHNOLOGY Outstanding Paper Award in 2011, the ENI Award Debut in
[29] A. U. Zgraggen, L. Fagiano, and M. Morari, “Real-time optimization and Research Prize in 2010, the Maffezzoni Prize in 2009, and a Marie Curie
adaptation of the crosswind flight of tethered wings for airborne wind International Outgoing Fellowship. He is an Associate Editor of the
energy,” IEEE Trans. Control Syst. Technol., vol. 23, no. 2, pp. 434–448, IEEE T RANSACTIONS ON C ONTROL S YSTEMS T ECHNOLOGY.
Mar. 2015.
[30] L. Fagiano, K. Huynh, B. Bamieh, and M. Khammash, “On sensor fusion
for airborne wind energy systems,” IEEE Trans. Control Syst. Technol.,
vol. 22, no. 3, pp. 930–943, May 2014.
[31] H. K. Khalil, Nonlinear Systems, 3rd ed. Englewood Cliffs, NJ, USA:
Prentice-Hall, 2001.
[32] F. Amato, Robust Control of Linear Systems Subject to Uncertain
Time-Varying Parameters (Lecture Notes in Control and Information
Sciences), vol. 325. Berlin, Germany: Springer-Verlag, 2006.
[33] L. Fagiano, M. Milanese, and D. Piga, “Optimization of airborne wind
energy generators,” Int. J. Robust Nonlinear Control, vol. 22, no. 18,
pp. 2055–2083, Dec. 2012.
[34] Swiss Kite Power. (Oct. 2013). Experimental Test Movie. [Online].
Available: http://youtu.be/yDRc3Ze4GAM Manfred Morari (F’05) received the Diploma
[35] I. Argatov, P. Rautakorpi, and R. Silvennoinen, “Estimation of the degree in chemical engineering from ETH Zurich,
mechanical energy output of the kite wind generator,” Renew. Energy, Zurich, Switzerland, the Ph.D. degree in chemical
vol. 34, no. 6, pp. 1525–1532, 2009. engineering from the University of Minnesota,
[36] P. Williams, B. Lansdorp, and W. Ockels, “Nonlinear control and esti- Minneapolis, MN, USA, and the Doctor Honoris
mation of a tethered kite in changing wind conditions,” J. Guid., Control Causa degree from Babes-Bolyai University,
Dyn., vol. 31, no. 3, pp. 793–799, May 2008. [Online]. Available: Cluj-Napoca, Romania.
http://dx.doi.org/10.2514/1.31604 He was the McCollum-Corcoran Professor of
Chemical Engineering and an Executive Officer for
Control and Dynamical Systems with the California
Institute of Technology, Pasadena, CA, USA.
He was the Head of the Automatic Control Laboratory from 1994 to 2008.
Aldo U. Zgraggen received the B.S. and He was the Head of the Department of Information Technology and Electrical
M.S. degrees in mechanical engineering from Engineering with ETH Zurich from 2009 to 2012. He has held appointments
ETH in 2007 and 2009, respectively, and the with Exxon, Irving, TX, USA, and ICI plc., London, U.K. His current
Ph.D. degree in information technology and research interests include hybrid systems and the control of biomedical
electrical engineering from the Automatic Control systems.
Laboratory, ETH Zurich, in 2014. Prof. Morari is a fellow of the International Federation of Automatic
He was with the Control and Dynamical Systems Control and the American Institute of Chemical Engineers (AIChE). In
Group, California Institute of Technology, Pasadena, recognition of his research contributions, he received numerous awards, such
CA, USA, from 2008 to 2009. In 2012, he was a as the Donald P. Eckman Award, the John R. Ragazzini Award and the
Visiting Student with the University of California at Richard E. Bellman Control Heritage Award of the American Automatic
Santa Barbara, Santa Barbara, CA, USA. His current Control Council, the Allan P. Colburn Award and the Professional Progress
research interests include control of tethered wings for airborne wind energy Award of AIChE, the Curtis W. McGraw Research Award of the American
generation. Society for Engineering Education, and the IEEE Control Systems Technical
Dr. Zgraggen was on the Program Committee of the Airborne Wind Energy Field Award. He was elected to the National Academy of Engineering (U.S.).
Conference in 2015. He serves on the technical advisory boards of several major corporations.

Authorized licensed use limited to: Consortium - Algeria (CERIST). Downloaded on June 08,2023 at 05:20:22 UTC from IEEE Xplore. Restrictions apply.

You might also like