SlideShare a Scribd company logo
1 of 12
Download to read offline
Determination of kite forces using three-dimensional flight trajectories
for ship propulsion
George M. Dadd*, Dominic A. Hudson, R.A. Shenoi
University of Southampton, Southampton, England SO17 1BJ, United Kingdom
a r t i c l e i n f o
Article history:
Received 15 November 2010
Accepted 23 January 2011
Available online 21 April 2011
Keywords:
Kite
Dynamics
Trajectories
Ship propulsion
Optimisaton
Experiment
a b s t r a c t
For application of kites to ships for power and propulsion, a scheme for predicting time averaged kite
forces is required. This paper presents a method for parameterizing figure of eight shape kite trajec-
tories and for predicting kite velocity, force and other performance characteristics. Results are
presented for a variety of maneuver shapes, assuming realistic performance characteristics from an
experimental test kite. Using a 300 m2
kite, with 300 m long flying lines in 6.18 msÀ1
wind, a time
averaged propulsive force of 16.7 tonne is achievable. A typical kite force polar is presented and
a sensitivity study is carried out to identify the importance of various parameters in the ship kite
propulsion system. Small horizontally orientated figure of eights shape kite trajectories centred on an
elevation of 15
is preferred for maximizing propulsive benefit. Propulsive force is found to be highly
sensitive to aspect ratio. Increasing aspect ratio from 4 to 5 is estimated to yield up to 15% more
drive force.
Ó 2011 Published by Elsevier Ltd.
1. Introduction
Kite propulsion is an attractive means to reduce fuel
consumption on ships by assisting the main engine using the power
of the wind. Recent developments, such as in autopilot kite control
and in launch and recovery systems1
have enabled them to be used
commercially for trans-oceanic voyages, yielding financial savings
through reduced fuel costs as well as minimizing emissions that are
harmful to the environment.
The determination of drive forces using a kite performance
model is required for ship velocity prediction, for enabling design,
for synthesising fuel savings and for optimizing kite systems for the
best propulsive effect. In addition, a kite performance model can be
used to implement carefully considered kite trajectories for
a desired force output.
Kite performance prediction models have been previously
established by Lloyd, [1], Wellicome [2], Naaijen [3], Williams [4]
and Argatov [5,6] although only Wellicome’s zero mass theory
has received published experimental validation. Dadd et al. (2010)
previously used the zero mass kite manoeuvring theory [2] to
predict kite line tension and other performance parameters. These
results were compared with real kite trajectories that had been
recorded using a purpose-specific kite dynamometer. The results
were shown to agree favourably; that work focused on the
validation of performance prediction based on kite position only.
The onset velocity and resulting line tension were calculated
without directly knowing the kite velocity itself. This paper focuses
on the additional modelling required in order to determine kite
velocity theoretically, an essential feature to enable the kite
performance to be established as a function of time.
Section 2 in this paper discusses the assumptions made in the
kite performance model. Section 3 presents a method for creating
kite trajectory shapes theoretically [2] and extends previous
developments by allowing the parameterized kite trajectories to be
transformed to simulate different mean angles to the wind. Section
4 defines the mathematical model. Section 5 describes the imple-
mentation and presents results using a case study for a typical ship
kite propulsion system. A new kite force polar diagram is developed
showing the propulsive drive for different wind angles. The
investigations are carried out considering the influence of the
Earth’s natural boundary layer. Section 6 presents an optimization
and sensitivity study that shows how various parameters effect
system performance including elevation, kite aspect ratio, angle of
attack, maneuver pole separation and pole circle size. Section 7
provides validation by way of comparison between theoretical
and experimental results [7].
* Corresponding author. Tel.: þ44 7815 044873.
E-mail address: georgedadd@hotmail.com (G.M. Dadd).
1
Pamphlet “Skysails Technological Information” available from www.skysails.
com.
Contents lists available at ScienceDirect
Renewable Energy
journal homepage: www.elsevier.com/locate/renene
0960-1481/$ e see front matter Ó 2011 Published by Elsevier Ltd.
doi:10.1016/j.renene.2011.01.027
Renewable Energy 36 (2011) 2667e2678
2. Assumptions in the kite force model
1. The zero mass theory [2] assumes that the kite and the lines are
weightless. This is reasonable provided that the real weight is
very small compared to the aerodynamic forces, as shown by
Dadd et al. [7].
2. The kite is assumed to maneuver on the surface of a sphere of
radius defined by the flying line.
3. The kite lift and drag coefficients are assumed to remain
constant. The aerodynamic lift and drag coefficients are given
by expressions of the form
CL ¼ f ðae; RnÞ (1)
and
CD ¼ gðae; RnÞ: (2)
This implies firstly that the dependence of the force coefficients on
Rn is negligible and secondly that the angle of attack is unchanging.
To explain the third assumption, the kite is in a condition of
force equilibrium during static flight, where the line tension is
equally opposed to the aerodynamic force (neglecting weight). The
kite assumes its position in the flight envelope where this condition
is met and the effective angle of attack (ae) is dependent on the
relative wind velocity and the angle of mount to the flying lines.
During dynamic flight, the kite seeks the same force equilibrium
condition. When an imbalance of force arises, the kite accelerates
almost instantaneously to achieve the apparent kite onset velocity
at which this equilibrium is again achieved. The angle of attack
remains the same as the static flight case where CL and CD are
constant.
The Reynolds number effects which can also influence CL and CD
are not expressly included in the zero mass model, although it is
noted that from Dadd et al. [7] that Rn was seen to vary between
7 Â 105
for static flight and 4.3 Â 106
for dynamic flight using a small
3 m2
kite in light winds. These are above the critical Rn number
(w5 Â105
) at which transition between laminar to turbulent flow
tends to occur and thus it can be expected that the flow will remain
substantially turbulent during dynamic flight and expectedly more
so for larger kites or for stronger winds. Thus with transition
between laminar and turbulent flow being unlikely during normal
flying conditions, the Rn effects are very minor and safe to neglect
whilst maintaining good predictions for kite performance.
Based on the above principles, Wellicome showed that the onset
wind velocity at the kite can be established in terms of its azimuth
and elevation spherical position angles, using the fundamental zero
mass equation [2].
U ¼ VA
cos qcos f
sin 3
: (3)
Here, q ¼ 0 f ¼ 0 defines the downwind direction.
Lloyd [1] had found that where the kite passes directly through
the downwind position, the onset velocity can be approximated by
Nomenclature
AK projected kite area, m2
AR aspect ratio
e aerodynamic planform efficiency factor (lifting line
theory)
f,g generic functions
F aerodynamic force magnitude, N
CL lift coefficient
CLa
lift coefficient at a
CL0
lift coefficient at a ¼ 0
CD drag coefficient
CD0
drag coefficient at a ¼ 0
D drag force magnitude, N
E rotation matrix
H pole of trajectory sphere
l aerodynamic lift force unit vector, N
L lift force magnitude, N
n exponent dependant on atmospheric and surface
conditions
n vector normal to great circle (right to left sweeps), m
n1,n2,n3 components of vector n, m
m vector normal to great circle (left to right sweeps), m
m1,m2,m3 components of vector m, m
O origin of trajectory sphere
P pole of small circle sweep
Q pole of small circle sweep
r kite position unit vector, m
ro small circle pole position vector, m
R kite position vector magnitude, m
R kite position vector, m
Re Reynolds number (Uck/n)
T time taken to traverse between two maneuver points A
and B, s
u onset velocity unit vector, msÀ1
U intersection node on trajectory
U onset velocity vector, msÀ1
U onset velocity magnitude, msÀ1
v apparent wind velocity unit vector, msÀ1
V intersection node on trajectory
V apparent wind velocity magnitude at the kite when
static, msÀ1
VT true wind speed, msÀ1
VTref
true wind speed at reference altitude, msÀ1
W intersection node on trajectory
V apparent wind at the kite, as though it were static,
msÀ1
x,y,z Cartesian position coordinates, m
X,Y,Z global Cartesian position coordinates, m
dt time step, s
f, g generic functions
ae effective angle of attack, 
a semi-vertex cone angle, 
a1 semi-vertex cone angle at P, 
a2 semi-vertex cone angle at Q, 
b azimuth angle of air onset velocity, 
d variable, 
3 aerodynamic drag angle, 
g elevation angle of air onset velocity, 
f azimuth angle, 
h1,2,3 transformation rotation angles about axis X, Y and Z, 
q elevation angle, 
ra density of air (1.19 kg mÀ3
at 20, 1 bar)
s variable, 
s variable, 
m substitution variable, (m ¼ 1/2raAKCLsec3)
z variable, 
G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e26782668
the apparent wind speed multiplied by the lift to drag ratio of the
kite. Thus, with present day kites, such as the Flexifoil blade II para-
foil type test kite, used by Dadd [7], having lift to drag ratios of the
order of six, onset velocities six times wind speed are typical for
downwind dynamic kite flight and since aerodynamic force
increases with the square of the onset velocity, line tensions thirty
six times that of static kite flight are feasible.
Practical kite trajectories must remain downwind of the kite
tether. The preferred maneuver shapes are those that do not result in
undesirable twisting of the flying lines or collision with the ground.
Consequently, vertical or horizontal figure of eight type motions are
commonplace for substantially downwind sailing courses, whilst
sine-wave (up and down) motions and static kite flight are
commonplace for upwind sailing courses. The manner in which
propulsive forces develop is largely dependent on the shape of
the kite trajectory adopted. Section 3 describes the way in which the
trajectories in the present paper were parameterized mathematically.
3. Parameterization of kite trajectory shapes
The trajectory parameterization follows the work of Wellicome
[2] and is further modified to allow vertically and diagonally
orientated trajectory shapes. The figure of eight flying maneuver
shape can be mathematically defined as comprising two small
circle sweeps for the ends of the maneuver and two great circle
sweeps which connect the small circular ends (see Figs. 1 and 2).
The great circle sweeps are along the intersections of a plane
passing through the origin with the sphere surface. The small circle
ends are defined using a semi-vertex angle for the cone swept out
by the radius and the position of the circle pole at its centre.
The spherical position angles, azimuth and elevation, are as
shown in Fig. 1. The position vector of the kite is given by
r ¼ (cos q cos f, cos qsin f, sin q). The unit vector n ¼ (1,2,3) is
chosen normal to the plane of the great circle sweep such that
nr ¼ 0.
rn1cos q cos f þ n2cos q sin f þ n3sin q ¼ 0 (4)
rtan q ¼ À

n1
n2
cos f þ
n2
n3
sin f

(5)
Hence Eq. (5) can be used to establish q, 4 ordinates on any great
circle defined by n.
The direction of the kite is defined by the value of f

=q

which can be determined by differentiation of Eq. (4) for
great circle sweeps as
ð À n1cos q sin f þ n2cos q cos fÞf

þ ð À n1sin q cos f À n2sin q sin f þ n3cos qÞq

¼ 0:
(6)
r
f

q
 ¼
sin qðn1cos f þ n2sin fÞ À n3cos q
cos qðn2cos f À n1sin fÞ
: (7)
For the small circle ends in Fig. 2, the pole P with position vector ro
defines the centre of the small circle sweep that is inscribed by the
kite with semi-vertex angle a. Points which lie on the small circular
arc conform to
r  ro ¼ cos a; (8)
where
r ¼ ðcos q cos f; cos q sin f; sin qÞ (9)
and
ro ¼ ðcos qocos fo; cos qosin fo; sin qoÞ: (10)
rcos a ¼ cos q cos qocosðf À foÞ þ sin q sin qo; (11)
rf ¼ fo Æ cosÀ1

cos a À sin q sin qo
cos q cos qo

: (12)
Hence Eq. (12) can be used to establish q, f ordinates on any small
circle sweep. The direction of flight for the small circle ends is
determined through differentiation of Eq. (11) as
f

q
 ¼
cos q sin qo À sin q cos qocosðf À foÞ
cos q cos qosinðf À foÞ
: (13)
Thus, to fully define the maneuver geometry, two poles (P(q1, f1)
and Q(q2, f2)) are required with corresponding semi-vertex angles
(a1 and a2) for each of the small circle ends, as well as calculation of
the two normal vectors n and m to define the great circle sweeps,
right to left and left to right respectively. It is necessary to establish
these maneuver segments as forming a continuous path by
ensuring tangency at the respective joins. For this, consideration is
given to the spherical triangle shown in Fig. 3, with a view to
finding the ratios n1/n3 and n2/n3 which can later be used to ensure
that tangency and continuity are correctly obtained. PH and QH in
Fig. 3 are great circles passing through the Z axis and PQ is another
great circle passing through O. The angle subtended between OP
and OH is 90 À q1, the angle subtended between OQ and OH is
90 À q2, the angle subtended between OP and OQ is z and the angle
subtended between HP and HQ is f1 À f2.
Applying the standard spherical cosine formula to triangle HPQ,
z can be found.
Fig. 1. Great circle sweep geometry.
Fig. 2. Small circle sweep geometry.
G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e2678 2669
cos z ¼ cosð90 À q1Þcosð90 À q2Þ
þ sinð90 À q1Þsinð90 À q2Þcosðf1 À f2Þ (14)
or
cosz ¼ sinq1sinq2 þ cosq1cosq2cosðf1 À f2Þ (15)
Fig. 4 shows the intersection node points for consecutive great
circle and small circle sweeps at R, T, U and V. z1 and z2 are the
angles subtended between OP and OS and between OQ and OS
respectively. a1, a2, s1 and s2 equivalently represent their included
angles with the origin. The angle formed by the intersection of PQ
and RT is d. The angle RPQ is s1, the angle HOR is 90 À qR, the angle
HOP is 90 À q1 and the angle RHP is f1 À 4R. HR, HP, RP, PQ, RT and
QT are all great circles. DPRS and DSTQ are right triangles.
Applying the spherical sine formula to DPRS,
sin a1
sin d
¼
sin z1
sin 90
(16)
rsin d ¼
sin a1
sin z1
(17)
Similarly for DSQT,
sin d ¼
sin a2
sin z2
(18)
Through equating Eqs. (17) and (18) and using z2 ¼ z À z1, the
value of z1 can be found,
tan z1 ¼
sin z sin a1
sin a2 þ cos z sin a1
(19)
Applying the cosine formula to DPRS, s1 can be obtained with
cos z1 ¼ cos a1cos s1 þsin a1sin s1cos 90 ¼ cos a1cos s1 (20)
rcos s1 ¼
cos z1
cos a1
(21)
Applying sine formula to DPRS, s1 is obtained using
sin s1 ¼
sin s1
sin z1
(22)
Employing the sine formula with DHPQ in Fig. 3 gives
sin

HbPQ

¼
cos q2sinðf1 À f2Þ
sin z
(23)
then
HbPQ ¼ HbPQ À s1: (24)
Applying cosine formula to DHPR enables calculation of qR,
cosð90 À qRÞ ¼ cos a1cosð90 À q1Þ
þ sin a1sinð90 À q1Þcos HbPQ (25)
rsinqR ¼ cosa1sinq1 þ sina1cosq1cos

HbPQ

(26)
Then, applying the sine formula to DHPR allows fR to be solved
using
sinðf1 À fRÞ ¼
sin a1sin

HbPQ

cos qR
(27)
The use of Eqs. (15)e(27) allows the locations the nodes qR and fR to
be solved for a given input set [q1, f1, a1, q2, f2, a2] that define the
maneuver shape. A similar set of equations can be developed to
establish the values of qT and fT. These values for points R and T can
be input in Eq. (4) and solved to establish the ratios n1/n2 and n3/n2.
Thus Eq. (5) may be used to define a series of data points along the
great circle sweep, RT. A similar process is repeated to establish
a series of positions along the other great circle sweep, UV. Points
which lie on the small circular end about the pole P are established
using Eq. (12) and similarly for points which lie on the small
circular end about pole Q. Data points on the surface of the sphere
which form a continuous and tangent curve around a figure of eight
shape have now been defined.
The parameterized trajectory may be transformed using Euler
rotations to achieve a vertically or diagonally orientated figure of
eight maneuver shape positioned at different angles relative to the
onset wind. Rotations are applied to the position vector, r, first about
the X axis, then about Y and about Z respectively in that order as
r0
¼ E3E2E1r: (28)
The rotation angles h1, h2 and h3 about X, Y and Z respectively
are user selected and the rotation matrices are E1, E2 and E3
given byFig. 4. Spherical triangles PRS, QST and HPR.
Fig. 3. Spherical triangle HPQ.
G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e26782670
E1 ¼
2
4
1 0 0
0 cosh1 sinh1
0 Àsinh1 cosh1
3
5;E2 ¼
2
4
cosh2 0 Àsinh2
0 1 0
sinh2 0 cosh2
3
5andE3
¼
2
4
cosh3 sinh3 0
Àsinh3 cosh3 0
0 0 1
3
5 (29)
f

=q

is then determined numerically in place of Eqs. (7) and (13).
4. Determination of kite forces, kite velocity and time
averaged quantities
The remainder of this section essentially follows the develop-
ment of Wellicome [2]. It is intended here to provide the necessary
background to Section 5.
The onset velocity can be expressed as
U ¼ Uu ¼ VAv À Rr

(30)
The aerodynamic force magnitude is given by
F ¼ L sec 3 ¼
1
2
raAKU2
CLsec 3 ¼ mU2
(31)
The Lift force magnitude is given by
L ¼
1
2
raAKU2
CL ¼ mU2
cos 3 (32)
The drag force magnitude is given by
D ¼
1
2
raAKU2
CD ¼ mU2
sin 3 (33)
By vector addition
F ¼ Fr ¼ Ll þ Du (34)
where l is a unit vector perpendicular to u so that lu ¼ 0.
Substituting for FL and D,
mU2
r ¼ mU2
cos 3l þ mU2
sin 3u (35)
or
r ¼ l cos 3 þ u sin 3 (36)
Taking the scalar product with onset unit direction u,
r  u ¼ l  u cos e þ u  u sin 3
rr  u ¼ sin 3
(37)
By taking scalar products with the position unit vector, Eq. (30)
becomes:
Uu  r ¼ U sin 3 ¼ Vv$r À R_r  r (38)
Since kite motion is confined to the surface of a sphere, its
velocity is tangential to the spheres’ surface such that ðr$r

¼ 0Þ. It
follows that
U ¼
Vv$r
sin 3
(39)
A Cartesian right handed system is chosen with the wind parallel to
the X axis. The position angles of the vectors R and U are as shown
in Fig. 5. The position angles of U are b and g. The unit vectors v, r
and u are expressed in terms of their position angles as
v ¼ f1; 0; 0g (40)
r ¼ fcos q cos f; cos q sin f; sin qg (41)
u ¼ fcos g cos b; cos g sin b; sin gg (42)
r

is then obtained by differentiation of r as
r

¼
n
À cos q sin ff

À sin q cos fq

; cos q cos ff

À sin q sin f_q; cos q_q
o
(43)
Substituting the expression for r

into 15 yields
Ufcos g cos b; cos g sin b; sin gg
¼
È
V þ R cos q sin ff

þ R sin q cos f q

; R sin q sin fq

À R cos q cos ff

; ÀR cos q q
 É
(44)
Since r  v ¼ cos q cos f, Eq. (39) can be expressed as
U ¼ VA
cos q cos f
sin 3
(45)
which is the same as Eq. (3), thus the onset velocity of air onto the
kite can be calculated according to any given kite position in terms
of azimuth and elevation, such as those positions which may be
defined using the parameterization of kite trajectory shapes
developed in Section 3 or alternatively, by using experimentally
recorded trajectories.
In real conditions, the wind speed can be expected to increase
with altitude due to viscous effects that slow the air in the Earth’s
atmospheric boundary layer. Calculation of Eq. (45) can therefore
be improved by employing [8]
VT ¼ VTref
Z
Zref
!n
(46)
to account for the difference in altitude relative to the wind refer-
ence height. The value of the exponent n may be expected to vary
depending on atmospheric conditions, however a value of 1/7 is
suitable for typical conditions at sea [8]. The lift force is calculated
using
L ¼
1
2
rAKU2
CL: (47)
The drag force is calculated using
Fig. 5. Flight envelope showing position angles for the kite and the onset velocity, U.
G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e2678 2671
D ¼
1
2
rAKU2
CD: (48)
The total aerodynamic kite force, assumed here to be equal to
the line tension is obtained as
F ¼
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
L2 þ D2
p
: (49)
or
F ¼
1
2
rAU2
CF (50)
The established onset velocity in Eq. (45) does not require or allow
the calculation of the kite velocity. This is obtained through the
implementation of the following formulae, Eqs. (51)e(58). Then, it
becomes possible to integrate the kite force around the maneuver
in order to determine time averaged force, magnitude and direc-
tion, using Eqs. (59)e(63).
The X, Yand Z components of Eq. (44) can be written separately as
X : U cos g cos b ¼ V þ cos q sin f Rf

þ sin q cos f Rq

(51)
Y : U cos g sin b ¼ Àcos q cos fRf

þ sin q sin fRq

(52)
Z : U sin g ¼ Àcos qR_q (53)
Combining Eqs. (53) with (45) yields an expression for R_q, for the
velocity of the kite in the q direction.
rRq

¼ ÀV
cos f sin g
sin 3
(54)
Substituting Eqs. (45) and (54) in Eq. (52) yields an expression
for R _f, for the velocity of the kite in the f direction,
Rf

¼ ÀV
tan q sin f sin g þ cos g sin b
sin 3
!
: (55)
Eqs. (45) and (54) can be combined to yield an expression for the
direction of the kite from kite flight dynamic considerations,
f

q
 ¼
tan q sin fsin g þ cos gsin b
cos fsin g
(56)
Eq. (51) combined with Eqs. (45), (54) and (55) and after manipu-
lation becomes
sin 3 ¼ sin q sin g þ cos q cos g cosðf À bÞ; (57)
or
b ¼ f Æ cosÀ1 sin 3 À sin q sin g
cos q cos g
!
(58)
For a stationary kite, g ¼ b ¼ 0, so that substitution of these values
into Eq. (58) enables determination of azimuth and elevation values
along the line of static equilibrium that limits the motion of the kite.
The time taken for the kite to maneuver from a point qA to qB can
be obtained from
T ¼
ZqB
qA
dt
dq
dq ¼
ZqB
qA
1
q
 dq (59)
and the time averaged aerodynamic force F is given by
F ¼
1
T
Z
Fdt ¼
1
T
ZqB
qA
F
dt
dq
dq ¼
1
T
ZqB
qA
F
q
 dq (60)
The angle of elevation for the time averaged force F is deter-
mined as
q ¼ tanÀ1
0
B
@
Fy
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
F2
x þ F2
y
r
1
C
A (61)
The azimuth angle for the time averaged force is determined as
f ¼ tanÀ1

Fy
Fx

(62)
The magnitude of force acting in the horizontal plane for useful
ship propulsion is thus obtained as
Fxy ¼ F cos q (63)
5. Implementation and results
Using the procedure outlined in Section 3, a horizontally
orientated figure of eight maneuver was defined as the default
maneuver shape (q1 ¼ À25, f1 ¼7, a1 ¼8, q2 ¼ 25, f2 ¼ 7, a2 ¼ 8).
This was manipulated using Euler rotation angles h1, h2 and h3 in Eq.
(29) to produce vertically orientated trajectories with 15 different
mean wind angles across the flight envelope. The Euler angles for
transformations are shown in Table 1 and the resulting maneuver
shapes are shown graphically in Fig. 6.
The theory of Sections 3 and 4 was implemented using Matlab
[9] to obtain performance results. The flow diagram in Fig. 7 shows
the order in which calculations were done.
The kite performance input parameters were previously
measured for an experimental test kite (Model: Flexifoil Blade III
3 m2
) similar to the type of kite that is typically used for ship
propulsion. These data are shown in Table 2 from Dadd et al. [7]. For
this purely theoretical study, the kite area used is 300 m2
and
default line length used is 300 m. These values are in keeping with
typical kite sizes and line lengths that are presently used on
commercially available ship kite propulsion systems.
During the implementation, it was found that for some input
sets, more than one solution was feasible for the parameters b and
g. The correct solution selected was that which ensured continuous
Table 1
Euler rotation angles for vertical trajectory manipulation.
Trajectory h1 h2 h3
1 105 35 67
2 102 35 60
3 100 35 50
4 98 35 40
5 96 35 30
6 94 35 20
7 92 35 10
8 90 35 0
9 88 35 À10
10 86 35 À20
11 84 35 À30
12 82 35 À40
13 80 35 À50
14 78 35 À60
15 75 35 À67
G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e26782672
results throughout the maneuver trajectory and satisfied the
requirement that the proper sense of the force vector was achieved.
Selected results arising for trajectory number 8 (highlighted in
Fig. 6) are shown in are in Fig. 8. These show the azimuth and
elevation angles that define kite position as an input, the kite
velocity together with the air onset velocity and the aerodynamic
force generated. These results show how the kite trajectory shape is
related to force output and provides a basis with which shapes can
be tailored to attain the desired output. This may be useful for
producing a control algorithm to regulate the force output. In
Fig. 8B it can be noted that the kite velocity closely follows the onset
velocity. This shows that the majority of the onset velocity is due to
the motion of the kite itself and explains why large force amplifi-
cation arises as a result of traversing the kite across the sky. Fig. 8C
shows that the line tension reduces significantly from 2 Â 105
to
0.3 Â 105
N as it is elevated during the maneuver. This is expected
since the kite is moved away from its location of maximum force
output which occurs directly downwind.
For predicting ship performance, it is useful to provide a single
time averaged force vector for a particular maneuver shape. The
time averaged aerodynamic forces were obtained using Eqs. (59)
and (60) for each of the trajectories in Table 1. To demonstrate
the effect of the wind increase with altitude due to the Earths
atmospheric boundary layer, kite performance was investigated
first using a wind gradient according to Eq. (46), and secondly with
no velocity gradient (see Fig. 9). The two cases have equal wind
speed at 10 m altitude (6.18 msÀ1
).
Fig.10A, shows the time averaged force for trajectories in Table 1,
not including the effect of the wind gradient, whilst Fig. 10B shows
the time averaged force in the presence of a wind gradient
(6.18 msÀ1
at Zref ¼ 10 m). The difference between the two
demonstrates the improved performance that arises as a result of
increasing wind speed with altitude. This is a significant added
benefit which is not realized by other wind propulsion systems,
such as conventional sails.
Fig. 10C shows the aerodynamic force which is realized
instantaneously as the kite is traversed along a straight path from
left to right at constant elevation, q ¼ 15.0 (inc. Wind gradient).
This result can be viewed alternatively as the force realised from
a figure of eight trajectory centred at q ¼ 15 with infinitely long
lines, such that the full extent of the trajectory may be considered
to remain at a single discreet value of azimuth and elevation. By
comparing Fig. 10B with Fig. 10C it is possible to see the extent by
which the propulsive drive force is reduced by traversing the kite
around the trajectories of Fig. 6. Fig. 10D shows the time averaged
aerodynamic forces from horizontally orientated trajectories that
have resulted from the optimization study in Section 6. These are
the maneuver shapes which are recommended for practical ship
propulsion.
6. Optimization and sensitivity study
Having defined a generic kite force polar diagram using a time
averaged result in both an even wind and in a wind gradient,
investigations have been undertaken to establish factors which
influence optimal trajectories. These factors include
1. The ideal elevation for maximizing line tension,
2. The ideal elevation to maximize the useful propulsive drive
force,
3. Kite aspect ratio,
Fig. 6. Vertical trajectory parameterization.
Fig. 7. Flow diagram for implementation.
Table 2
statically measured performance characteristics for the test kite (Model: Flexifoil
Blade III 3 m).
3 L=D CF CL CD
N samples 21 21 21 21 21
Mean 9.55 6.07 0.786 0.776 0.128
Standard deviation 1.48 0.88 0.115 0.116 0.012
95% Confidence limit 0.63 0.38 0.049 0.050 0.005
99% Confidence limit 0.83 0.50 0.065 0.065 0.006
G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e2678 2673
4. Angle of attack,
5. Maneuver pole separation and Pole circle size.
6.1. Elevation for maximizing line tension
The wind speed increases with altitude in the Earth’s natural
boundary layer. Hence the onset velocity in Eq. (45) and the
resulting lift through Eq. (47) is also increased. However, the
elevation angle in Eq. (45) reduces the onset velocity due to its
influence on kite motions. There is an optimum elevation at which
the line tension is maximized. In order to find this Eqs. (45) and (46)
are combined with Z ¼ R sin q to obtain onset velocity (U) as
a function of its kite position angles,
U ¼ VTref
R sin q
Zref
!n
cos q cos f
sin 3
: (64)
The maximum onset velocity (and thus lift) with respect to
elevation is the solution to DU/Dq ¼ 0, subject to D2
U=Dq2
 0,
hence
n
R sin q
Zref
!nÀ1
R cos2q
Zref
À sin q
R sin q
Zref
!n
¼ 0 (65)
Taking n ¼ 1/7 for typical conditions at sea [8], Zref ¼ 10 m,
R ¼ 300 m and 3 ¼ 9.55, the solution to Eq. (65) is numerically
obtained as q ¼ 20.7. Therefore in the presence of a wind gradient
caused by the Earths boundary layer, the onset velocity and hence
line tension can be maximized by maintaining a trajectory that is
Fig. 8. Parameters arising through trajectory number 8. (A) Position angles azimuth and
elevation; (B) kite onset flow velocity and actual kite velocity; (C) aerodynamic force.
Fig. 9. Wind profiles.
Fig. 10. Horizontal component of kite force, FAxy
[N/104
] for different cases. (A) Time
averaged vertical trajectories, wind speed ¼ 6.18 msÀ1
, wind gradient off. (B) Time
averaged vertical trajectories, wind speed ¼ 6.18 msÀ1
at Zref ¼ 10 m, wind gradient on.
(C) Instantaneous kite force along straight horizontal trajectory q ¼ 15, wind
speed ¼ 6.18 msÀ1
at Zref ¼ 10 m, wind gradient on. (D) Time averaged horizontal
trajectories, wind speed 6.18 msÀ1
at Zref ¼ 10 m, wind gradient on.
G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e26782674
centred closely to this value of elevation. This result is useful where
the absolute value of the line tension magnitude is important, such
as for producing electricity using kite power.
6.2. Elevation for maximizing propulsive drive force
For ship propulsion, only the horizontal component of the line
tension leads to useful propulsive force. The horizontal component
reduces with the cosine of elevation, so that the optimum elevation
is obtained by maximizing the value of U cos q.
This is given by
U cos q ¼ VTref
R sin q
Zref
!n
cos2q cos f
sin 3
(66)
Setting D(U cos q)/Dq ¼ 0, subject to D2
U cos q=Dq2
 0 to find
the maximum gives,
n
R sin q
Zref
!nÀ1
R
Zref
!
$cos2
q À 2 sin q
R sin q
Zref
!n
¼ 0 (67)
When solved numerically, Eq. (67) yields the maximum horizontal
component of drive force where q ¼ 15.0. Hence for ship pro-
pulsion in the presence of a wind gradient, an optimal trajectory
should centre closely to this elevation. In this respect, horizontally
orientated figure of eight trajectories are preferential to vertical
ones for downwind trajectories, since elevation angles during the
maneuver remain closer to the optimum value. However, for
courses sailed close to the wind a vertically orientated trajectory
must be used in order to maintain a close winded time averaged
force, such as for trajectories 1 and 15 in Table 1.
Trajectories which reflect this optimization have been investi-
gated using the default parameters (q1 ¼ À25, f1 ¼7, a1 ¼8, q2 ¼ 25,
f2 ¼ 7, a2 ¼ 8). These have been transformed using Eq. (28) and
Table 3. The resulting trajectory shapes are shown in Fig. 11. The
time averaged horizontal force output for each trajectory is shown
in Fig. 10D. By comparing Fig. 10B with Fig. 10D, horizontally
orientated maneuvers centred proximally to the optimum eleva-
tion are seen to make a significant improvement to the drive force
compared to vertical ones.
6.3. Kite aspect ratio
To investigate the effects of kite aspect ratio on aerodynamic
force, the kite drag angle is expressed as a function of the lift to drag
ratio using
3 ¼ tanÀ1

1
ðL=DÞ

(68)
Eqs. (45) and (50) are combined to yield an expression that relates
the drag angle to aerodynamic force as
FA ¼
1
2
rA

VAcos q cos f
sin 3
2
CF (69)
The drag coefficient of a kite with a dissimilar aspect ratio to that
used in the present study can be calculated using
C0
D ¼ CD þ
C2
L
p

1
AR0 À
1
AR

; (70)
where C0
D corresponds tothe drag coefficientof awingof a newaspect
ratio AR0 [10]. It is assumed here that the kite is set so that the lift
coefficient remains unchanged by modified aspect ratio. The lift to
drag ratio of a wing with modified aspect ratio is therefore given by
L
D
¼
CL
CD þ
C2
L
p

1
AR0 À
1
AR
: (71)
To find the sensitivity of the kite system to aspect ratio, the
derivative dFA=dðARÞ is sought. This is given by
dFA
dðARÞ
¼
dFA
d3
$
d3
dðL=DÞ
$
dðL=DÞ
dðARÞ
; (72)
for which Eqs. (68)e(71) are differentiated to obtain
dFA
d3
¼ ÀrACFðVAcos q cos fÞ
2cos 3
sin3
3
; (73)
d3
dðL=DÞ
¼ À
1
ðL=DÞ2
þ1
; (74)
and
dðL=DÞ
dðARÞ
¼
C3
L
pAR2
h
CD þ
C2
L
p

1
AR0 À
1
AR
#2
(75)
where 3 is in radians. Using r ¼ 1.19, A ¼ 300, VA ¼ 6.18, q ¼ 15,
f ¼ 0, CL ¼ 0.776, CD ¼ 0.128, CF ¼ 0.786 and AR ¼ 4.86, Table 4 was
formulated using Eq. (68)e(75), for a range of values of aspect ratio
modified from the experimentally measured performance data in
Table 2.
Table 3
Euler rotation angles for horizontal trajectory manipulation.
Trajectory h1 h2 h3
1 0 15 À48
2 0 15 À40
3 0 15 À32
4 0 15 À24
5 0 15 À16
6 0 15 À8
7 0 15 0
8 0 15 8
9 0 15 16
10 0 15 24
11 0 15 32
12 0 15 40
13 0 15 48
Fig. 11. Horizontal trajectory parameterization.
G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e2678 2675
For the example input set, Table 4 shows that increasing aspect
ratio from 4 to 5 for example, results in a dramatically improved
instantaneousaerodynamicforce.19.21 Â104
N is achieved insteadof
16.67 Â 104
N yielding 15% more aerodynamic force per unit increase
in aspect ratio. There are other factors that require consideration in
selectionofkiteaspect ratio,suchashandling characteristics,stability
and response ingusts.Anecdotal evidence[11]suggeststhat theseare
factors which would tend to favour lower aspect ratios.
6.4. Kite angle of attack
The lift coefficient at which the kite operates has a significant
influence on aerodynamic force; firstly through its direct influence
on the lift generated, but also through a secondary effect due to the
change in the operational L/D that accompanies a change in angle of
attack. The optimum angle of attack for maximizing aerodynamic
force may be determined by considering Prandtl lifting line theory to
establish a functional relationship that can be assessed analytically
when combined with the zero mass theory. If an increase in lift
coefficientoccurstogetherwith anincrease in liftto drag,thepositive
effect of each will be compounded. If an increase in lift coefficient
accompanies a reduction in lift to drag, the effect will be negated. The
optimum condition for maximized propulsive force occurs when
dFA
da
¼
dFA
dCL
$
dCL
da
þ
dFA
dðL=DÞ
$
dðL=DÞ
da
¼ 0: (76)
From Eqs. (50), (45) and (68)
dFA
dCL
¼
1
2
rAU2
¼
1
2
rA

VA
cos q cos f
sin 3
2
; (77)
and through combining (50), (68) and (71),
dFA
dðL=DÞ
¼

À rACFðVAcos q cos fÞ
2cos 3
sin3
3
'
$
(
À
1
ðL=DÞ2
þ1
)
:
(78)
The 3D lift coefficient from Prandtl lifting line theory is
CLa
¼ CL0
þ 2p

AR
AR þ 2

a: (79)
The induced drag coefficient from Prandtl lifting line theory is
CD ¼ CD0
þ
C2
L
pARe
: (80)
From Eq. (79)
dCL
da
¼ 2p

AR
AR þ 2

: (81)
Combining Eqs. (79) and (80) gives
L=D ¼
CL0
þ 2p
À AR
AR þ 2
Á
a
CD0
þ
À
CL0
þ 2p
À AR
AR þ 2
Á
a
Á2
pARe
(82)
Differentiating Eq. (82) gives
Table 4
Effect of aspect ratio on aerodynamic force.
AR’ 3 [
] L/D FA [104
N] d(FA)/d(AR) [104
N]
1 19.86 2.77 4.34 5.25
2 13.37 4.21 9.36 4.60
3 11.11 5.09 13.46 3.62
4 9.97 5.69 16.67 2.84
5 9.29 6.12 19.21 2.26
6 8.83 6.44 21.24 1.83
7 8.50 6.69 22.91 1.51
8 8.25 6.90 24.29 1.27
9 8.06 7.06 25.46 1.08
10 7.90 7.20 26.46 0.92
dðL=DÞ
da
¼
0
@CD0
þ
À
CL0
þ 2p
À AR
AR þ 2
Á
a
Á2
pARe
1
A$2p

AR
AR þ 2

À

CL0
þ 2p

AR
AR þ 2

a

$
0
@
2CL0
$2p
À AR
AR þ 2
Á
pARe
þ
8p2
À AR
AR þ 2
Á2
a
pARe
1
A
0
@CD0
þ
À
CL0
þ 2p
À AR
AR þ 2
Á
a
Á2
pARe
1
A
2
(83)
Fig. 12. Trajectories shapes with different pole circle sizes.
Fig. 13. Trajectory shapes with different pole circle separation.
G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e26782676
Using AR ¼ 4.86, CD0
¼ 0:05, CL0
¼ 0:337, ra ¼ 1.19, A ¼ 300,
V ¼ 6:18, q ¼ 15, f ¼ 0, and e ¼ 0.9 Eq. (76) was found yield
a ¼ 15.28. This is the theoretical optimum as determined by
Prandtl lifting line theory, though there are other factors which
might influence the best angle of attack for practical use. In
particular, the way that the kite recovers from off-design flight
conditions, such as stall, is likely to be significantly improved by
mounting the kite to operate at a smaller angle of attack.
6.5. Pole separation and circle size
The effects of maneuver pole separation and pole circle size
were investigated by comparing a series of different kite trajecto-
ries (See Figs. 12 and 13). Results were obtained through imple-
mentation of the flow chart in Fig. 7, including the effect of the wind
gradient as determined by Eq. (46). The effect of increasing pole
circle size and pole circle separation on time averaged force output
is shown in Figs. 14 and 15 respectively. These figures show that
increasing pole circle size and separation both lead to a marked
reduction of time averaged propulsive drive force since the
maneuver is moved away from its optimal position for maximized
propulsive effect. The limitation of these investigations is that no
account is taken of the potential reduction in lift coefficient asso-
ciated with canopy deformation and angle of roll during a tight
turn.
7. Experimental validation
Previously Eq. (45), for obtaining onset velocity as a function of
kite position, was validated by comparing predicted aerodynamic
force with measured kite line tension [7]. The kite forces, force
direction and wind speed had been recorded by flying a kite from
a dynamometer in natural winds. In the present paper, the
implementation of the theory has been developed to allow the kite
velocity and motions to be computationally predicted, as well as
the instantaneous force based on position. For experimental
validation of the present work, two trajectories are shown in
Fig. 16.
The theoretically parameterized trajectory is defined according
to Section 3, whilst the experimentally determined trajectory was
recorded using a purpose built kite dynamometer described by
Dadd [7]. The theoretically determined shape was selected to
closely match that of the experimental one (see Fig. 16). Small
differences are apparent since the manually flown trajectory
exhibits irregularities. The aerodynamic force was determined in
three ways (see Fig. 17). First, the zero mass theoretical aero-
dynamic force was determined using the experimentally deter-
mined position coordinates as an input. Second, the actual
aerodynamic kite force was obtained using the dynamometer.
Comparing these two results serves to validate Eq. (45), which
determines kite onset velocity as a function of position, as
described by Dadd et al [7]. The third result in Fig. 16 is also based
on Eq. (45), but uses Eqs. (51)e(59) to establish the kite velocity
and hence to theoretically determine the results to a base of time,
by integrating around the maneuver. It can be seen that the
theoretical time for the kite to progress around the maneuver
closely matches that of the experimental kite trajectory. This
serves to validate Eqs. (51)e(59). The errors arising through
determination of kite force through each of these methods is
Fig. 14. Force reduction with circle size.
Fig. 15. Force reduction with pole separation.
Fig. 16. Experimentally and theoretically defined trajectories for comparison.
Fig. 17. Comparison between theoretical predictions and experimental measurements.
G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e2678 2677
discussed in detail by Dadd et al [7]. The most significant error for
experimentally determined kite trajectories was shown to be due
to spatial and temporal differences in wind speed across the flight
envelope which was not properly accounted for.
8. Conclusions
This paper has presented a scheme for predicting time averaged
kite forces together with a method for prameterizing figure of eight
shape kite trajectories. The theoretical model was modified to allow
a high degree of control over input maneuver paths in order to
assess the effects of maneuver orientation, shape and true wind
angle. The influence of the changing wind strength in the Earth’s
natural boundary layer is included. The scheme has been imple-
mented to show variation of onset velocity, kite velocity, onset
angles and aerodynamic force for an example maneuver. In a case
study with a 300 m2
kite, 300 m long flying lines and 6.18 msÀ1
wind speed (at 10 m), peak drive force up to 30 Â 104
N was pre-
dicted. A typical vertical figure of eight maneuver in the same
conditions was shown to produce a time averaged horizontal drive
force component 16.7 Â 104
N. A complete force polar showing the
horizontal component of aerodynamic force for different true wind
angles has thus been obtained.
The calculation of kite velocity and the propagation of the kite
around a theoretically defined maneuver have been validated by
comparing a real recorded trajectory to a simulated one. The time
scales for a theoretical and practical trajectory are shown to agree
favourably, as do the magnitudes of minimum and maximum kite
force. The increase in wind with altitude is found to double the aero-
dynamic force compared to a hypotheticalcase with nowind gradient.
Through considering the wind gradient in combination with kite
motions, the optimum mean kite elevation angle for ship pro-
pulsion during dynamic flight was found to be 15. Horizontally
orientated trajectories are shown to improve downwind perfor-
mance by a factor of 1.4 compared to vertical ones, since the
maneuver can remain closer to the optimum elevation angle for
propulsion. Vertical maneuvers are, however, preferred when
maintaining close winded performance is necessary.
By combining Prandtl lifting line theory with the zero mass
model, the derivative dFA=dðARÞ was calculated at a variety of
different aspect ratios. For the downwind dynamic flight condition,
increasing the kite aspect ratio from 4 to 5, for example, was found
to yield a 15% increase in drive force showing that force is highly
sensitive to aspect ratio through its effect on kite L/D. Low aspect
ratios may however be favourable to kite stability characteristics.
The angle of attack at which the kite operates modifies the
aerodynamic force generated through its influence on lift and drag
coefficients, but also through its effect on L/D. Optimizing the angle
of attack with respect to both of these yields a ¼ 15.
Optimal trajectories were found to be those that are centred
most closely to the best region of the flight envelope. Increasing
maneuver pole circle size and pole separation both diminished the
time averaged propulsive drive force arising. If the potential effect
of reduced CL during tight turns is negligible, the best maneuver is
found to be a figure of eight shape with small circular ends and
small pole separation, centred closely to an elevation of 15.
References
[1] Loyd ML. Crosswind kite power (for large-scale wind power production). AIAA
Journal of Energy 1980;4(3):106e11.
[2] Wellicome, JF, Wilkinson S. Propulsive kites e an initial study. University of
Southampton, Department of Ship Science, Faculty of Engineering and Applied
Science. Tech. Rept. SSSU 19; 1984.
[3] Naaijen P, Koster V, Dallinga RP. On the power savings by an auxiliary kite
propulsion system. International Shipbuilding Progress 2006;53(4):255e79.
[4] Williams P. Optimal wind power extraction with a tethered kite. AIAA
Guidance, Navigation, and Control Conference and Exhibit. AIAA, Keystone,
Colorado; 2006.
[5] Argatov I, Rautakorpi P, Silvennoinen R. Estimation of the mechanical energy
output of the kite wind generator. Renewable Energy 2009;34(6):1525e32.
[6] Argatov I, Silvennoinen R. Energy conversion efficiency of the pumping kite
wind generator. Renewable Energy 2010;35(5):1052e60.
[7] Dadd GM, Hudson DA, Shenoi RA. Comparison of two kite force models with
experiment. Journal of Aircraft 2010;47(1):212e24.
[8] Claughton AR, Shenoi RA, Wellicome JF. Sailing yacht design theory. 1 ed.
Harlow: Addison Wesley Longman; 1998. pp. 5e6.
[9] Matlab. Mathmatics laboratory, software package [Ver. 7.0.0]. Cambridge:
MathWorks; 2004.
[10] Abbott IH, Doenhoff AE. Theory of wing sections. New York: Dover Publica-
tions; 1959.
[11] Bryant, LW, Brown WS. Collected researches on the stability of kites and
towed gliders. National Phisics Laboratory, Reports and Memoranda 2303;
1942.
G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e26782678

More Related Content

What's hot

Simulation of Deployment and Operation of an Earth Observing Satellite
Simulation of Deployment and Operation of an Earth Observing SatelliteSimulation of Deployment and Operation of an Earth Observing Satellite
Simulation of Deployment and Operation of an Earth Observing Satellite
Alber Douglawi
 
“The strong attraction of gravity at Geoscience Australia” (Richard Lane, Geo...
“The strong attraction of gravity at Geoscience Australia” (Richard Lane, Geo...“The strong attraction of gravity at Geoscience Australia” (Richard Lane, Geo...
“The strong attraction of gravity at Geoscience Australia” (Richard Lane, Geo...
Richard Lane
 

What's hot (18)

Simulation of Deployment and Operation of an Earth Observing Satellite
Simulation of Deployment and Operation of an Earth Observing SatelliteSimulation of Deployment and Operation of an Earth Observing Satellite
Simulation of Deployment and Operation of an Earth Observing Satellite
 
Poirel1997
Poirel1997Poirel1997
Poirel1997
 
WING_accepted
WING_acceptedWING_accepted
WING_accepted
 
Optimal trajectory to Saturn in ion-thruster powered spacecraft
Optimal trajectory to Saturn in ion-thruster powered spacecraftOptimal trajectory to Saturn in ion-thruster powered spacecraft
Optimal trajectory to Saturn in ion-thruster powered spacecraft
 
PosterForAGU
PosterForAGUPosterForAGU
PosterForAGU
 
Computational and Experimental Study of the Aerodynamics of a Micro Air Vehicle
Computational and Experimental Study of the Aerodynamics of a Micro Air VehicleComputational and Experimental Study of the Aerodynamics of a Micro Air Vehicle
Computational and Experimental Study of the Aerodynamics of a Micro Air Vehicle
 
Simulation of Wind Power Dynamic for Electricity Production in Nassiriyah Dis...
Simulation of Wind Power Dynamic for Electricity Production in Nassiriyah Dis...Simulation of Wind Power Dynamic for Electricity Production in Nassiriyah Dis...
Simulation of Wind Power Dynamic for Electricity Production in Nassiriyah Dis...
 
2009 Dransfield - conforming falcon gravity and the global gravity anomaly
2009 Dransfield - conforming falcon gravity and the global gravity anomaly2009 Dransfield - conforming falcon gravity and the global gravity anomaly
2009 Dransfield - conforming falcon gravity and the global gravity anomaly
 
Numerical Modelling of Wind Patterns around a Solar Parabolic Trough Collector
Numerical Modelling of Wind Patterns around a Solar Parabolic Trough CollectorNumerical Modelling of Wind Patterns around a Solar Parabolic Trough Collector
Numerical Modelling of Wind Patterns around a Solar Parabolic Trough Collector
 
Optimisation of the design of uav wing j.alexander, Prakash, BSM Augustine
Optimisation of the design of uav wing   j.alexander, Prakash, BSM AugustineOptimisation of the design of uav wing   j.alexander, Prakash, BSM Augustine
Optimisation of the design of uav wing j.alexander, Prakash, BSM Augustine
 
International Journal of Engineering and Science Invention (IJESI)
International Journal of Engineering and Science Invention (IJESI)International Journal of Engineering and Science Invention (IJESI)
International Journal of Engineering and Science Invention (IJESI)
 
“The strong attraction of gravity at Geoscience Australia” (Richard Lane, Geo...
“The strong attraction of gravity at Geoscience Australia” (Richard Lane, Geo...“The strong attraction of gravity at Geoscience Australia” (Richard Lane, Geo...
“The strong attraction of gravity at Geoscience Australia” (Richard Lane, Geo...
 
garretttylerreport3
garretttylerreport3garretttylerreport3
garretttylerreport3
 
EAGES Proceedings - D. N. Sinitsyn
EAGES Proceedings - D. N. SinitsynEAGES Proceedings - D. N. Sinitsyn
EAGES Proceedings - D. N. Sinitsyn
 
Impact of Different Wake Models on the Estimation of Wind Farm Power Generation
Impact of Different Wake Models on the Estimation of Wind Farm Power GenerationImpact of Different Wake Models on the Estimation of Wind Farm Power Generation
Impact of Different Wake Models on the Estimation of Wind Farm Power Generation
 
CFD Analysis Of Savonius Vertical Axis Wind Turbine: A Review
CFD Analysis Of Savonius Vertical Axis Wind Turbine: A ReviewCFD Analysis Of Savonius Vertical Axis Wind Turbine: A Review
CFD Analysis Of Savonius Vertical Axis Wind Turbine: A Review
 
Thesis Summary
Thesis SummaryThesis Summary
Thesis Summary
 
Implementation of Wilcox k − ω model to UCNS3D
Implementation of Wilcox k − ω model to UCNS3DImplementation of Wilcox k − ω model to UCNS3D
Implementation of Wilcox k − ω model to UCNS3D
 

Viewers also liked (6)

Daniel Stone CV January 2016
Daniel Stone CV January 2016Daniel Stone CV January 2016
Daniel Stone CV January 2016
 
IRFAN.RESUME UPDTD
IRFAN.RESUME UPDTDIRFAN.RESUME UPDTD
IRFAN.RESUME UPDTD
 
Educación
EducaciónEducación
Educación
 
Po co komu psychometria?
Po co komu psychometria?Po co komu psychometria?
Po co komu psychometria?
 
Manish- Sales Manager
Manish- Sales ManagerManish- Sales Manager
Manish- Sales Manager
 
Hoja de calculo
Hoja de calculoHoja de calculo
Hoja de calculo
 

Similar to Determination of Kite forces for ship propulsion

On The Form Factor Prediction Of A Displacement Type Vessel: JBC Case
On The Form Factor Prediction Of A Displacement Type Vessel: JBC CaseOn The Form Factor Prediction Of A Displacement Type Vessel: JBC Case
On The Form Factor Prediction Of A Displacement Type Vessel: JBC Case
Ismail Topal
 
NACA 4412 Lab Report Final
NACA 4412 Lab Report FinalNACA 4412 Lab Report Final
NACA 4412 Lab Report Final
Gregory Day
 
Optimisation of the design of uav wing j.alexander
Optimisation of the design of uav wing   j.alexanderOptimisation of the design of uav wing   j.alexander
Optimisation of the design of uav wing j.alexander
sathyabama
 
Aerodynamic Characteristics of Dragonfly Wing Sections
Aerodynamic Characteristics of Dragonfly Wing SectionsAerodynamic Characteristics of Dragonfly Wing Sections
Aerodynamic Characteristics of Dragonfly Wing Sections
Pradeep Kannah Purushothaman
 

Similar to Determination of Kite forces for ship propulsion (20)

IRJET-Subsonic Flow Study and Analysis on Rotating Cylinder Airfoil
IRJET-Subsonic Flow Study and Analysis on Rotating Cylinder AirfoilIRJET-Subsonic Flow Study and Analysis on Rotating Cylinder Airfoil
IRJET-Subsonic Flow Study and Analysis on Rotating Cylinder Airfoil
 
Effect of Gap between Airfoil and Embedded Rotating Cylinder on the Airfoil A...
Effect of Gap between Airfoil and Embedded Rotating Cylinder on the Airfoil A...Effect of Gap between Airfoil and Embedded Rotating Cylinder on the Airfoil A...
Effect of Gap between Airfoil and Embedded Rotating Cylinder on the Airfoil A...
 
On The Form Factor Prediction Of A Displacement Type Vessel: JBC Case
On The Form Factor Prediction Of A Displacement Type Vessel: JBC CaseOn The Form Factor Prediction Of A Displacement Type Vessel: JBC Case
On The Form Factor Prediction Of A Displacement Type Vessel: JBC Case
 
Full paper jbc icame2016
Full paper jbc icame2016Full paper jbc icame2016
Full paper jbc icame2016
 
C04571829
C04571829C04571829
C04571829
 
NACA 4412 Lab Report Final
NACA 4412 Lab Report FinalNACA 4412 Lab Report Final
NACA 4412 Lab Report Final
 
dighe (3)
dighe (3)dighe (3)
dighe (3)
 
Optimisation of the design of uav wing j.alexander
Optimisation of the design of uav wing   j.alexanderOptimisation of the design of uav wing   j.alexander
Optimisation of the design of uav wing j.alexander
 
Analysis Of Owl-Like Airfoil Aerodynamics At Low Reynolds Number Flow
Analysis Of Owl-Like Airfoil Aerodynamics At Low Reynolds Number FlowAnalysis Of Owl-Like Airfoil Aerodynamics At Low Reynolds Number Flow
Analysis Of Owl-Like Airfoil Aerodynamics At Low Reynolds Number Flow
 
Dynamic aerodynamic structural coupling numerical simulation on the flexible ...
Dynamic aerodynamic structural coupling numerical simulation on the flexible ...Dynamic aerodynamic structural coupling numerical simulation on the flexible ...
Dynamic aerodynamic structural coupling numerical simulation on the flexible ...
 
Determination of Flutter Angle by Resolving Effective Gyroscope Couple to Ret...
Determination of Flutter Angle by Resolving Effective Gyroscope Couple to Ret...Determination of Flutter Angle by Resolving Effective Gyroscope Couple to Ret...
Determination of Flutter Angle by Resolving Effective Gyroscope Couple to Ret...
 
Alterations of the Tip Vortex Structure from a Hovering Rotor using Passive T...
Alterations of the Tip Vortex Structure from a Hovering Rotor using Passive T...Alterations of the Tip Vortex Structure from a Hovering Rotor using Passive T...
Alterations of the Tip Vortex Structure from a Hovering Rotor using Passive T...
 
Wind Energy to Electrical Energy
Wind Energy to  Electrical EnergyWind Energy to  Electrical Energy
Wind Energy to Electrical Energy
 
CFD analysis of bio-inspired corrugated airfoils
CFD analysis of bio-inspired corrugated airfoilsCFD analysis of bio-inspired corrugated airfoils
CFD analysis of bio-inspired corrugated airfoils
 
Aerodynamic Characteristics of Dragonfly Wing Sections
Aerodynamic Characteristics of Dragonfly Wing SectionsAerodynamic Characteristics of Dragonfly Wing Sections
Aerodynamic Characteristics of Dragonfly Wing Sections
 
Analysis of wings using Airfoil NACA 4412 at different angle of attack
Analysis of wings using Airfoil NACA 4412 at different angle of attackAnalysis of wings using Airfoil NACA 4412 at different angle of attack
Analysis of wings using Airfoil NACA 4412 at different angle of attack
 
Prediction of aerodynamic characteristics for slender bluff bodies with nose ...
Prediction of aerodynamic characteristics for slender bluff bodies with nose ...Prediction of aerodynamic characteristics for slender bluff bodies with nose ...
Prediction of aerodynamic characteristics for slender bluff bodies with nose ...
 
Aerodynamics Of Insect S Flight
Aerodynamics Of Insect S FlightAerodynamics Of Insect S Flight
Aerodynamics Of Insect S Flight
 
CFD Analysis of a Three Bladed H-Rotor of Vertical Axis Wind Turbine
CFD Analysis of a Three Bladed H-Rotor of   Vertical Axis Wind Turbine CFD Analysis of a Three Bladed H-Rotor of   Vertical Axis Wind Turbine
CFD Analysis of a Three Bladed H-Rotor of Vertical Axis Wind Turbine
 
Characteristics of shock reflection in the dual solution domain
Characteristics of shock reflection in the dual solution domainCharacteristics of shock reflection in the dual solution domain
Characteristics of shock reflection in the dual solution domain
 

Determination of Kite forces for ship propulsion

  • 1. Determination of kite forces using three-dimensional flight trajectories for ship propulsion George M. Dadd*, Dominic A. Hudson, R.A. Shenoi University of Southampton, Southampton, England SO17 1BJ, United Kingdom a r t i c l e i n f o Article history: Received 15 November 2010 Accepted 23 January 2011 Available online 21 April 2011 Keywords: Kite Dynamics Trajectories Ship propulsion Optimisaton Experiment a b s t r a c t For application of kites to ships for power and propulsion, a scheme for predicting time averaged kite forces is required. This paper presents a method for parameterizing figure of eight shape kite trajec- tories and for predicting kite velocity, force and other performance characteristics. Results are presented for a variety of maneuver shapes, assuming realistic performance characteristics from an experimental test kite. Using a 300 m2 kite, with 300 m long flying lines in 6.18 msÀ1 wind, a time averaged propulsive force of 16.7 tonne is achievable. A typical kite force polar is presented and a sensitivity study is carried out to identify the importance of various parameters in the ship kite propulsion system. Small horizontally orientated figure of eights shape kite trajectories centred on an elevation of 15 is preferred for maximizing propulsive benefit. Propulsive force is found to be highly sensitive to aspect ratio. Increasing aspect ratio from 4 to 5 is estimated to yield up to 15% more drive force. Ó 2011 Published by Elsevier Ltd. 1. Introduction Kite propulsion is an attractive means to reduce fuel consumption on ships by assisting the main engine using the power of the wind. Recent developments, such as in autopilot kite control and in launch and recovery systems1 have enabled them to be used commercially for trans-oceanic voyages, yielding financial savings through reduced fuel costs as well as minimizing emissions that are harmful to the environment. The determination of drive forces using a kite performance model is required for ship velocity prediction, for enabling design, for synthesising fuel savings and for optimizing kite systems for the best propulsive effect. In addition, a kite performance model can be used to implement carefully considered kite trajectories for a desired force output. Kite performance prediction models have been previously established by Lloyd, [1], Wellicome [2], Naaijen [3], Williams [4] and Argatov [5,6] although only Wellicome’s zero mass theory has received published experimental validation. Dadd et al. (2010) previously used the zero mass kite manoeuvring theory [2] to predict kite line tension and other performance parameters. These results were compared with real kite trajectories that had been recorded using a purpose-specific kite dynamometer. The results were shown to agree favourably; that work focused on the validation of performance prediction based on kite position only. The onset velocity and resulting line tension were calculated without directly knowing the kite velocity itself. This paper focuses on the additional modelling required in order to determine kite velocity theoretically, an essential feature to enable the kite performance to be established as a function of time. Section 2 in this paper discusses the assumptions made in the kite performance model. Section 3 presents a method for creating kite trajectory shapes theoretically [2] and extends previous developments by allowing the parameterized kite trajectories to be transformed to simulate different mean angles to the wind. Section 4 defines the mathematical model. Section 5 describes the imple- mentation and presents results using a case study for a typical ship kite propulsion system. A new kite force polar diagram is developed showing the propulsive drive for different wind angles. The investigations are carried out considering the influence of the Earth’s natural boundary layer. Section 6 presents an optimization and sensitivity study that shows how various parameters effect system performance including elevation, kite aspect ratio, angle of attack, maneuver pole separation and pole circle size. Section 7 provides validation by way of comparison between theoretical and experimental results [7]. * Corresponding author. Tel.: þ44 7815 044873. E-mail address: georgedadd@hotmail.com (G.M. Dadd). 1 Pamphlet “Skysails Technological Information” available from www.skysails. com. Contents lists available at ScienceDirect Renewable Energy journal homepage: www.elsevier.com/locate/renene 0960-1481/$ e see front matter Ó 2011 Published by Elsevier Ltd. doi:10.1016/j.renene.2011.01.027 Renewable Energy 36 (2011) 2667e2678
  • 2. 2. Assumptions in the kite force model 1. The zero mass theory [2] assumes that the kite and the lines are weightless. This is reasonable provided that the real weight is very small compared to the aerodynamic forces, as shown by Dadd et al. [7]. 2. The kite is assumed to maneuver on the surface of a sphere of radius defined by the flying line. 3. The kite lift and drag coefficients are assumed to remain constant. The aerodynamic lift and drag coefficients are given by expressions of the form CL ¼ f ðae; RnÞ (1) and CD ¼ gðae; RnÞ: (2) This implies firstly that the dependence of the force coefficients on Rn is negligible and secondly that the angle of attack is unchanging. To explain the third assumption, the kite is in a condition of force equilibrium during static flight, where the line tension is equally opposed to the aerodynamic force (neglecting weight). The kite assumes its position in the flight envelope where this condition is met and the effective angle of attack (ae) is dependent on the relative wind velocity and the angle of mount to the flying lines. During dynamic flight, the kite seeks the same force equilibrium condition. When an imbalance of force arises, the kite accelerates almost instantaneously to achieve the apparent kite onset velocity at which this equilibrium is again achieved. The angle of attack remains the same as the static flight case where CL and CD are constant. The Reynolds number effects which can also influence CL and CD are not expressly included in the zero mass model, although it is noted that from Dadd et al. [7] that Rn was seen to vary between 7 Â 105 for static flight and 4.3 Â 106 for dynamic flight using a small 3 m2 kite in light winds. These are above the critical Rn number (w5 Â105 ) at which transition between laminar to turbulent flow tends to occur and thus it can be expected that the flow will remain substantially turbulent during dynamic flight and expectedly more so for larger kites or for stronger winds. Thus with transition between laminar and turbulent flow being unlikely during normal flying conditions, the Rn effects are very minor and safe to neglect whilst maintaining good predictions for kite performance. Based on the above principles, Wellicome showed that the onset wind velocity at the kite can be established in terms of its azimuth and elevation spherical position angles, using the fundamental zero mass equation [2]. U ¼ VA cos qcos f sin 3 : (3) Here, q ¼ 0 f ¼ 0 defines the downwind direction. Lloyd [1] had found that where the kite passes directly through the downwind position, the onset velocity can be approximated by Nomenclature AK projected kite area, m2 AR aspect ratio e aerodynamic planform efficiency factor (lifting line theory) f,g generic functions F aerodynamic force magnitude, N CL lift coefficient CLa lift coefficient at a CL0 lift coefficient at a ¼ 0 CD drag coefficient CD0 drag coefficient at a ¼ 0 D drag force magnitude, N E rotation matrix H pole of trajectory sphere l aerodynamic lift force unit vector, N L lift force magnitude, N n exponent dependant on atmospheric and surface conditions n vector normal to great circle (right to left sweeps), m n1,n2,n3 components of vector n, m m vector normal to great circle (left to right sweeps), m m1,m2,m3 components of vector m, m O origin of trajectory sphere P pole of small circle sweep Q pole of small circle sweep r kite position unit vector, m ro small circle pole position vector, m R kite position vector magnitude, m R kite position vector, m Re Reynolds number (Uck/n) T time taken to traverse between two maneuver points A and B, s u onset velocity unit vector, msÀ1 U intersection node on trajectory U onset velocity vector, msÀ1 U onset velocity magnitude, msÀ1 v apparent wind velocity unit vector, msÀ1 V intersection node on trajectory V apparent wind velocity magnitude at the kite when static, msÀ1 VT true wind speed, msÀ1 VTref true wind speed at reference altitude, msÀ1 W intersection node on trajectory V apparent wind at the kite, as though it were static, msÀ1 x,y,z Cartesian position coordinates, m X,Y,Z global Cartesian position coordinates, m dt time step, s f, g generic functions ae effective angle of attack, a semi-vertex cone angle, a1 semi-vertex cone angle at P, a2 semi-vertex cone angle at Q, b azimuth angle of air onset velocity, d variable, 3 aerodynamic drag angle, g elevation angle of air onset velocity, f azimuth angle, h1,2,3 transformation rotation angles about axis X, Y and Z, q elevation angle, ra density of air (1.19 kg mÀ3 at 20, 1 bar) s variable, s variable, m substitution variable, (m ¼ 1/2raAKCLsec3) z variable, G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e26782668
  • 3. the apparent wind speed multiplied by the lift to drag ratio of the kite. Thus, with present day kites, such as the Flexifoil blade II para- foil type test kite, used by Dadd [7], having lift to drag ratios of the order of six, onset velocities six times wind speed are typical for downwind dynamic kite flight and since aerodynamic force increases with the square of the onset velocity, line tensions thirty six times that of static kite flight are feasible. Practical kite trajectories must remain downwind of the kite tether. The preferred maneuver shapes are those that do not result in undesirable twisting of the flying lines or collision with the ground. Consequently, vertical or horizontal figure of eight type motions are commonplace for substantially downwind sailing courses, whilst sine-wave (up and down) motions and static kite flight are commonplace for upwind sailing courses. The manner in which propulsive forces develop is largely dependent on the shape of the kite trajectory adopted. Section 3 describes the way in which the trajectories in the present paper were parameterized mathematically. 3. Parameterization of kite trajectory shapes The trajectory parameterization follows the work of Wellicome [2] and is further modified to allow vertically and diagonally orientated trajectory shapes. The figure of eight flying maneuver shape can be mathematically defined as comprising two small circle sweeps for the ends of the maneuver and two great circle sweeps which connect the small circular ends (see Figs. 1 and 2). The great circle sweeps are along the intersections of a plane passing through the origin with the sphere surface. The small circle ends are defined using a semi-vertex angle for the cone swept out by the radius and the position of the circle pole at its centre. The spherical position angles, azimuth and elevation, are as shown in Fig. 1. The position vector of the kite is given by r ¼ (cos q cos f, cos qsin f, sin q). The unit vector n ¼ (1,2,3) is chosen normal to the plane of the great circle sweep such that nr ¼ 0. rn1cos q cos f þ n2cos q sin f þ n3sin q ¼ 0 (4) rtan q ¼ À n1 n2 cos f þ n2 n3 sin f (5) Hence Eq. (5) can be used to establish q, 4 ordinates on any great circle defined by n. The direction of the kite is defined by the value of f =q which can be determined by differentiation of Eq. (4) for great circle sweeps as ð À n1cos q sin f þ n2cos q cos fÞf þ ð À n1sin q cos f À n2sin q sin f þ n3cos qÞq ¼ 0: (6) r f q ¼ sin qðn1cos f þ n2sin fÞ À n3cos q cos qðn2cos f À n1sin fÞ : (7) For the small circle ends in Fig. 2, the pole P with position vector ro defines the centre of the small circle sweep that is inscribed by the kite with semi-vertex angle a. Points which lie on the small circular arc conform to r ro ¼ cos a; (8) where r ¼ ðcos q cos f; cos q sin f; sin qÞ (9) and ro ¼ ðcos qocos fo; cos qosin fo; sin qoÞ: (10) rcos a ¼ cos q cos qocosðf À foÞ þ sin q sin qo; (11) rf ¼ fo Æ cosÀ1 cos a À sin q sin qo cos q cos qo : (12) Hence Eq. (12) can be used to establish q, f ordinates on any small circle sweep. The direction of flight for the small circle ends is determined through differentiation of Eq. (11) as f q ¼ cos q sin qo À sin q cos qocosðf À foÞ cos q cos qosinðf À foÞ : (13) Thus, to fully define the maneuver geometry, two poles (P(q1, f1) and Q(q2, f2)) are required with corresponding semi-vertex angles (a1 and a2) for each of the small circle ends, as well as calculation of the two normal vectors n and m to define the great circle sweeps, right to left and left to right respectively. It is necessary to establish these maneuver segments as forming a continuous path by ensuring tangency at the respective joins. For this, consideration is given to the spherical triangle shown in Fig. 3, with a view to finding the ratios n1/n3 and n2/n3 which can later be used to ensure that tangency and continuity are correctly obtained. PH and QH in Fig. 3 are great circles passing through the Z axis and PQ is another great circle passing through O. The angle subtended between OP and OH is 90 À q1, the angle subtended between OQ and OH is 90 À q2, the angle subtended between OP and OQ is z and the angle subtended between HP and HQ is f1 À f2. Applying the standard spherical cosine formula to triangle HPQ, z can be found. Fig. 1. Great circle sweep geometry. Fig. 2. Small circle sweep geometry. G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e2678 2669
  • 4. cos z ¼ cosð90 À q1Þcosð90 À q2Þ þ sinð90 À q1Þsinð90 À q2Þcosðf1 À f2Þ (14) or cosz ¼ sinq1sinq2 þ cosq1cosq2cosðf1 À f2Þ (15) Fig. 4 shows the intersection node points for consecutive great circle and small circle sweeps at R, T, U and V. z1 and z2 are the angles subtended between OP and OS and between OQ and OS respectively. a1, a2, s1 and s2 equivalently represent their included angles with the origin. The angle formed by the intersection of PQ and RT is d. The angle RPQ is s1, the angle HOR is 90 À qR, the angle HOP is 90 À q1 and the angle RHP is f1 À 4R. HR, HP, RP, PQ, RT and QT are all great circles. DPRS and DSTQ are right triangles. Applying the spherical sine formula to DPRS, sin a1 sin d ¼ sin z1 sin 90 (16) rsin d ¼ sin a1 sin z1 (17) Similarly for DSQT, sin d ¼ sin a2 sin z2 (18) Through equating Eqs. (17) and (18) and using z2 ¼ z À z1, the value of z1 can be found, tan z1 ¼ sin z sin a1 sin a2 þ cos z sin a1 (19) Applying the cosine formula to DPRS, s1 can be obtained with cos z1 ¼ cos a1cos s1 þsin a1sin s1cos 90 ¼ cos a1cos s1 (20) rcos s1 ¼ cos z1 cos a1 (21) Applying sine formula to DPRS, s1 is obtained using sin s1 ¼ sin s1 sin z1 (22) Employing the sine formula with DHPQ in Fig. 3 gives sin HbPQ ¼ cos q2sinðf1 À f2Þ sin z (23) then HbPQ ¼ HbPQ À s1: (24) Applying cosine formula to DHPR enables calculation of qR, cosð90 À qRÞ ¼ cos a1cosð90 À q1Þ þ sin a1sinð90 À q1Þcos HbPQ (25) rsinqR ¼ cosa1sinq1 þ sina1cosq1cos HbPQ (26) Then, applying the sine formula to DHPR allows fR to be solved using sinðf1 À fRÞ ¼ sin a1sin HbPQ cos qR (27) The use of Eqs. (15)e(27) allows the locations the nodes qR and fR to be solved for a given input set [q1, f1, a1, q2, f2, a2] that define the maneuver shape. A similar set of equations can be developed to establish the values of qT and fT. These values for points R and T can be input in Eq. (4) and solved to establish the ratios n1/n2 and n3/n2. Thus Eq. (5) may be used to define a series of data points along the great circle sweep, RT. A similar process is repeated to establish a series of positions along the other great circle sweep, UV. Points which lie on the small circular end about the pole P are established using Eq. (12) and similarly for points which lie on the small circular end about pole Q. Data points on the surface of the sphere which form a continuous and tangent curve around a figure of eight shape have now been defined. The parameterized trajectory may be transformed using Euler rotations to achieve a vertically or diagonally orientated figure of eight maneuver shape positioned at different angles relative to the onset wind. Rotations are applied to the position vector, r, first about the X axis, then about Y and about Z respectively in that order as r0 ¼ E3E2E1r: (28) The rotation angles h1, h2 and h3 about X, Y and Z respectively are user selected and the rotation matrices are E1, E2 and E3 given byFig. 4. Spherical triangles PRS, QST and HPR. Fig. 3. Spherical triangle HPQ. G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e26782670
  • 5. E1 ¼ 2 4 1 0 0 0 cosh1 sinh1 0 Àsinh1 cosh1 3 5;E2 ¼ 2 4 cosh2 0 Àsinh2 0 1 0 sinh2 0 cosh2 3 5andE3 ¼ 2 4 cosh3 sinh3 0 Àsinh3 cosh3 0 0 0 1 3 5 (29) f =q is then determined numerically in place of Eqs. (7) and (13). 4. Determination of kite forces, kite velocity and time averaged quantities The remainder of this section essentially follows the develop- ment of Wellicome [2]. It is intended here to provide the necessary background to Section 5. The onset velocity can be expressed as U ¼ Uu ¼ VAv À Rr (30) The aerodynamic force magnitude is given by F ¼ L sec 3 ¼ 1 2 raAKU2 CLsec 3 ¼ mU2 (31) The Lift force magnitude is given by L ¼ 1 2 raAKU2 CL ¼ mU2 cos 3 (32) The drag force magnitude is given by D ¼ 1 2 raAKU2 CD ¼ mU2 sin 3 (33) By vector addition F ¼ Fr ¼ Ll þ Du (34) where l is a unit vector perpendicular to u so that lu ¼ 0. Substituting for FL and D, mU2 r ¼ mU2 cos 3l þ mU2 sin 3u (35) or r ¼ l cos 3 þ u sin 3 (36) Taking the scalar product with onset unit direction u, r u ¼ l u cos e þ u u sin 3 rr u ¼ sin 3 (37) By taking scalar products with the position unit vector, Eq. (30) becomes: Uu r ¼ U sin 3 ¼ Vv$r À R_r r (38) Since kite motion is confined to the surface of a sphere, its velocity is tangential to the spheres’ surface such that ðr$r ¼ 0Þ. It follows that U ¼ Vv$r sin 3 (39) A Cartesian right handed system is chosen with the wind parallel to the X axis. The position angles of the vectors R and U are as shown in Fig. 5. The position angles of U are b and g. The unit vectors v, r and u are expressed in terms of their position angles as v ¼ f1; 0; 0g (40) r ¼ fcos q cos f; cos q sin f; sin qg (41) u ¼ fcos g cos b; cos g sin b; sin gg (42) r is then obtained by differentiation of r as r ¼ n À cos q sin ff À sin q cos fq ; cos q cos ff À sin q sin f_q; cos q_q o (43) Substituting the expression for r into 15 yields Ufcos g cos b; cos g sin b; sin gg ¼ È V þ R cos q sin ff þ R sin q cos f q ; R sin q sin fq À R cos q cos ff ; ÀR cos q q É (44) Since r v ¼ cos q cos f, Eq. (39) can be expressed as U ¼ VA cos q cos f sin 3 (45) which is the same as Eq. (3), thus the onset velocity of air onto the kite can be calculated according to any given kite position in terms of azimuth and elevation, such as those positions which may be defined using the parameterization of kite trajectory shapes developed in Section 3 or alternatively, by using experimentally recorded trajectories. In real conditions, the wind speed can be expected to increase with altitude due to viscous effects that slow the air in the Earth’s atmospheric boundary layer. Calculation of Eq. (45) can therefore be improved by employing [8] VT ¼ VTref Z Zref !n (46) to account for the difference in altitude relative to the wind refer- ence height. The value of the exponent n may be expected to vary depending on atmospheric conditions, however a value of 1/7 is suitable for typical conditions at sea [8]. The lift force is calculated using L ¼ 1 2 rAKU2 CL: (47) The drag force is calculated using Fig. 5. Flight envelope showing position angles for the kite and the onset velocity, U. G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e2678 2671
  • 6. D ¼ 1 2 rAKU2 CD: (48) The total aerodynamic kite force, assumed here to be equal to the line tension is obtained as F ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi L2 þ D2 p : (49) or F ¼ 1 2 rAU2 CF (50) The established onset velocity in Eq. (45) does not require or allow the calculation of the kite velocity. This is obtained through the implementation of the following formulae, Eqs. (51)e(58). Then, it becomes possible to integrate the kite force around the maneuver in order to determine time averaged force, magnitude and direc- tion, using Eqs. (59)e(63). The X, Yand Z components of Eq. (44) can be written separately as X : U cos g cos b ¼ V þ cos q sin f Rf þ sin q cos f Rq (51) Y : U cos g sin b ¼ Àcos q cos fRf þ sin q sin fRq (52) Z : U sin g ¼ Àcos qR_q (53) Combining Eqs. (53) with (45) yields an expression for R_q, for the velocity of the kite in the q direction. rRq ¼ ÀV cos f sin g sin 3 (54) Substituting Eqs. (45) and (54) in Eq. (52) yields an expression for R _f, for the velocity of the kite in the f direction, Rf ¼ ÀV tan q sin f sin g þ cos g sin b sin 3 ! : (55) Eqs. (45) and (54) can be combined to yield an expression for the direction of the kite from kite flight dynamic considerations, f q ¼ tan q sin fsin g þ cos gsin b cos fsin g (56) Eq. (51) combined with Eqs. (45), (54) and (55) and after manipu- lation becomes sin 3 ¼ sin q sin g þ cos q cos g cosðf À bÞ; (57) or b ¼ f Æ cosÀ1 sin 3 À sin q sin g cos q cos g ! (58) For a stationary kite, g ¼ b ¼ 0, so that substitution of these values into Eq. (58) enables determination of azimuth and elevation values along the line of static equilibrium that limits the motion of the kite. The time taken for the kite to maneuver from a point qA to qB can be obtained from T ¼ ZqB qA dt dq dq ¼ ZqB qA 1 q dq (59) and the time averaged aerodynamic force F is given by F ¼ 1 T Z Fdt ¼ 1 T ZqB qA F dt dq dq ¼ 1 T ZqB qA F q dq (60) The angle of elevation for the time averaged force F is deter- mined as q ¼ tanÀ1 0 B @ Fy ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi F2 x þ F2 y r 1 C A (61) The azimuth angle for the time averaged force is determined as f ¼ tanÀ1 Fy Fx (62) The magnitude of force acting in the horizontal plane for useful ship propulsion is thus obtained as Fxy ¼ F cos q (63) 5. Implementation and results Using the procedure outlined in Section 3, a horizontally orientated figure of eight maneuver was defined as the default maneuver shape (q1 ¼ À25, f1 ¼7, a1 ¼8, q2 ¼ 25, f2 ¼ 7, a2 ¼ 8). This was manipulated using Euler rotation angles h1, h2 and h3 in Eq. (29) to produce vertically orientated trajectories with 15 different mean wind angles across the flight envelope. The Euler angles for transformations are shown in Table 1 and the resulting maneuver shapes are shown graphically in Fig. 6. The theory of Sections 3 and 4 was implemented using Matlab [9] to obtain performance results. The flow diagram in Fig. 7 shows the order in which calculations were done. The kite performance input parameters were previously measured for an experimental test kite (Model: Flexifoil Blade III 3 m2 ) similar to the type of kite that is typically used for ship propulsion. These data are shown in Table 2 from Dadd et al. [7]. For this purely theoretical study, the kite area used is 300 m2 and default line length used is 300 m. These values are in keeping with typical kite sizes and line lengths that are presently used on commercially available ship kite propulsion systems. During the implementation, it was found that for some input sets, more than one solution was feasible for the parameters b and g. The correct solution selected was that which ensured continuous Table 1 Euler rotation angles for vertical trajectory manipulation. Trajectory h1 h2 h3 1 105 35 67 2 102 35 60 3 100 35 50 4 98 35 40 5 96 35 30 6 94 35 20 7 92 35 10 8 90 35 0 9 88 35 À10 10 86 35 À20 11 84 35 À30 12 82 35 À40 13 80 35 À50 14 78 35 À60 15 75 35 À67 G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e26782672
  • 7. results throughout the maneuver trajectory and satisfied the requirement that the proper sense of the force vector was achieved. Selected results arising for trajectory number 8 (highlighted in Fig. 6) are shown in are in Fig. 8. These show the azimuth and elevation angles that define kite position as an input, the kite velocity together with the air onset velocity and the aerodynamic force generated. These results show how the kite trajectory shape is related to force output and provides a basis with which shapes can be tailored to attain the desired output. This may be useful for producing a control algorithm to regulate the force output. In Fig. 8B it can be noted that the kite velocity closely follows the onset velocity. This shows that the majority of the onset velocity is due to the motion of the kite itself and explains why large force amplifi- cation arises as a result of traversing the kite across the sky. Fig. 8C shows that the line tension reduces significantly from 2 Â 105 to 0.3 Â 105 N as it is elevated during the maneuver. This is expected since the kite is moved away from its location of maximum force output which occurs directly downwind. For predicting ship performance, it is useful to provide a single time averaged force vector for a particular maneuver shape. The time averaged aerodynamic forces were obtained using Eqs. (59) and (60) for each of the trajectories in Table 1. To demonstrate the effect of the wind increase with altitude due to the Earths atmospheric boundary layer, kite performance was investigated first using a wind gradient according to Eq. (46), and secondly with no velocity gradient (see Fig. 9). The two cases have equal wind speed at 10 m altitude (6.18 msÀ1 ). Fig.10A, shows the time averaged force for trajectories in Table 1, not including the effect of the wind gradient, whilst Fig. 10B shows the time averaged force in the presence of a wind gradient (6.18 msÀ1 at Zref ¼ 10 m). The difference between the two demonstrates the improved performance that arises as a result of increasing wind speed with altitude. This is a significant added benefit which is not realized by other wind propulsion systems, such as conventional sails. Fig. 10C shows the aerodynamic force which is realized instantaneously as the kite is traversed along a straight path from left to right at constant elevation, q ¼ 15.0 (inc. Wind gradient). This result can be viewed alternatively as the force realised from a figure of eight trajectory centred at q ¼ 15 with infinitely long lines, such that the full extent of the trajectory may be considered to remain at a single discreet value of azimuth and elevation. By comparing Fig. 10B with Fig. 10C it is possible to see the extent by which the propulsive drive force is reduced by traversing the kite around the trajectories of Fig. 6. Fig. 10D shows the time averaged aerodynamic forces from horizontally orientated trajectories that have resulted from the optimization study in Section 6. These are the maneuver shapes which are recommended for practical ship propulsion. 6. Optimization and sensitivity study Having defined a generic kite force polar diagram using a time averaged result in both an even wind and in a wind gradient, investigations have been undertaken to establish factors which influence optimal trajectories. These factors include 1. The ideal elevation for maximizing line tension, 2. The ideal elevation to maximize the useful propulsive drive force, 3. Kite aspect ratio, Fig. 6. Vertical trajectory parameterization. Fig. 7. Flow diagram for implementation. Table 2 statically measured performance characteristics for the test kite (Model: Flexifoil Blade III 3 m). 3 L=D CF CL CD N samples 21 21 21 21 21 Mean 9.55 6.07 0.786 0.776 0.128 Standard deviation 1.48 0.88 0.115 0.116 0.012 95% Confidence limit 0.63 0.38 0.049 0.050 0.005 99% Confidence limit 0.83 0.50 0.065 0.065 0.006 G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e2678 2673
  • 8. 4. Angle of attack, 5. Maneuver pole separation and Pole circle size. 6.1. Elevation for maximizing line tension The wind speed increases with altitude in the Earth’s natural boundary layer. Hence the onset velocity in Eq. (45) and the resulting lift through Eq. (47) is also increased. However, the elevation angle in Eq. (45) reduces the onset velocity due to its influence on kite motions. There is an optimum elevation at which the line tension is maximized. In order to find this Eqs. (45) and (46) are combined with Z ¼ R sin q to obtain onset velocity (U) as a function of its kite position angles, U ¼ VTref R sin q Zref !n cos q cos f sin 3 : (64) The maximum onset velocity (and thus lift) with respect to elevation is the solution to DU/Dq ¼ 0, subject to D2 U=Dq2 0, hence n R sin q Zref !nÀ1 R cos2q Zref À sin q R sin q Zref !n ¼ 0 (65) Taking n ¼ 1/7 for typical conditions at sea [8], Zref ¼ 10 m, R ¼ 300 m and 3 ¼ 9.55, the solution to Eq. (65) is numerically obtained as q ¼ 20.7. Therefore in the presence of a wind gradient caused by the Earths boundary layer, the onset velocity and hence line tension can be maximized by maintaining a trajectory that is Fig. 8. Parameters arising through trajectory number 8. (A) Position angles azimuth and elevation; (B) kite onset flow velocity and actual kite velocity; (C) aerodynamic force. Fig. 9. Wind profiles. Fig. 10. Horizontal component of kite force, FAxy [N/104 ] for different cases. (A) Time averaged vertical trajectories, wind speed ¼ 6.18 msÀ1 , wind gradient off. (B) Time averaged vertical trajectories, wind speed ¼ 6.18 msÀ1 at Zref ¼ 10 m, wind gradient on. (C) Instantaneous kite force along straight horizontal trajectory q ¼ 15, wind speed ¼ 6.18 msÀ1 at Zref ¼ 10 m, wind gradient on. (D) Time averaged horizontal trajectories, wind speed 6.18 msÀ1 at Zref ¼ 10 m, wind gradient on. G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e26782674
  • 9. centred closely to this value of elevation. This result is useful where the absolute value of the line tension magnitude is important, such as for producing electricity using kite power. 6.2. Elevation for maximizing propulsive drive force For ship propulsion, only the horizontal component of the line tension leads to useful propulsive force. The horizontal component reduces with the cosine of elevation, so that the optimum elevation is obtained by maximizing the value of U cos q. This is given by U cos q ¼ VTref R sin q Zref !n cos2q cos f sin 3 (66) Setting D(U cos q)/Dq ¼ 0, subject to D2 U cos q=Dq2 0 to find the maximum gives, n R sin q Zref !nÀ1 R Zref ! $cos2 q À 2 sin q R sin q Zref !n ¼ 0 (67) When solved numerically, Eq. (67) yields the maximum horizontal component of drive force where q ¼ 15.0. Hence for ship pro- pulsion in the presence of a wind gradient, an optimal trajectory should centre closely to this elevation. In this respect, horizontally orientated figure of eight trajectories are preferential to vertical ones for downwind trajectories, since elevation angles during the maneuver remain closer to the optimum value. However, for courses sailed close to the wind a vertically orientated trajectory must be used in order to maintain a close winded time averaged force, such as for trajectories 1 and 15 in Table 1. Trajectories which reflect this optimization have been investi- gated using the default parameters (q1 ¼ À25, f1 ¼7, a1 ¼8, q2 ¼ 25, f2 ¼ 7, a2 ¼ 8). These have been transformed using Eq. (28) and Table 3. The resulting trajectory shapes are shown in Fig. 11. The time averaged horizontal force output for each trajectory is shown in Fig. 10D. By comparing Fig. 10B with Fig. 10D, horizontally orientated maneuvers centred proximally to the optimum eleva- tion are seen to make a significant improvement to the drive force compared to vertical ones. 6.3. Kite aspect ratio To investigate the effects of kite aspect ratio on aerodynamic force, the kite drag angle is expressed as a function of the lift to drag ratio using 3 ¼ tanÀ1 1 ðL=DÞ (68) Eqs. (45) and (50) are combined to yield an expression that relates the drag angle to aerodynamic force as FA ¼ 1 2 rA VAcos q cos f sin 3 2 CF (69) The drag coefficient of a kite with a dissimilar aspect ratio to that used in the present study can be calculated using C0 D ¼ CD þ C2 L p 1 AR0 À 1 AR ; (70) where C0 D corresponds tothe drag coefficientof awingof a newaspect ratio AR0 [10]. It is assumed here that the kite is set so that the lift coefficient remains unchanged by modified aspect ratio. The lift to drag ratio of a wing with modified aspect ratio is therefore given by L D ¼ CL CD þ C2 L p 1 AR0 À 1 AR : (71) To find the sensitivity of the kite system to aspect ratio, the derivative dFA=dðARÞ is sought. This is given by dFA dðARÞ ¼ dFA d3 $ d3 dðL=DÞ $ dðL=DÞ dðARÞ ; (72) for which Eqs. (68)e(71) are differentiated to obtain dFA d3 ¼ ÀrACFðVAcos q cos fÞ 2cos 3 sin3 3 ; (73) d3 dðL=DÞ ¼ À 1 ðL=DÞ2 þ1 ; (74) and dðL=DÞ dðARÞ ¼ C3 L pAR2 h CD þ C2 L p 1 AR0 À 1 AR #2 (75) where 3 is in radians. Using r ¼ 1.19, A ¼ 300, VA ¼ 6.18, q ¼ 15, f ¼ 0, CL ¼ 0.776, CD ¼ 0.128, CF ¼ 0.786 and AR ¼ 4.86, Table 4 was formulated using Eq. (68)e(75), for a range of values of aspect ratio modified from the experimentally measured performance data in Table 2. Table 3 Euler rotation angles for horizontal trajectory manipulation. Trajectory h1 h2 h3 1 0 15 À48 2 0 15 À40 3 0 15 À32 4 0 15 À24 5 0 15 À16 6 0 15 À8 7 0 15 0 8 0 15 8 9 0 15 16 10 0 15 24 11 0 15 32 12 0 15 40 13 0 15 48 Fig. 11. Horizontal trajectory parameterization. G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e2678 2675
  • 10. For the example input set, Table 4 shows that increasing aspect ratio from 4 to 5 for example, results in a dramatically improved instantaneousaerodynamicforce.19.21 Â104 N is achieved insteadof 16.67 Â 104 N yielding 15% more aerodynamic force per unit increase in aspect ratio. There are other factors that require consideration in selectionofkiteaspect ratio,suchashandling characteristics,stability and response ingusts.Anecdotal evidence[11]suggeststhat theseare factors which would tend to favour lower aspect ratios. 6.4. Kite angle of attack The lift coefficient at which the kite operates has a significant influence on aerodynamic force; firstly through its direct influence on the lift generated, but also through a secondary effect due to the change in the operational L/D that accompanies a change in angle of attack. The optimum angle of attack for maximizing aerodynamic force may be determined by considering Prandtl lifting line theory to establish a functional relationship that can be assessed analytically when combined with the zero mass theory. If an increase in lift coefficientoccurstogetherwith anincrease in liftto drag,thepositive effect of each will be compounded. If an increase in lift coefficient accompanies a reduction in lift to drag, the effect will be negated. The optimum condition for maximized propulsive force occurs when dFA da ¼ dFA dCL $ dCL da þ dFA dðL=DÞ $ dðL=DÞ da ¼ 0: (76) From Eqs. (50), (45) and (68) dFA dCL ¼ 1 2 rAU2 ¼ 1 2 rA VA cos q cos f sin 3 2 ; (77) and through combining (50), (68) and (71), dFA dðL=DÞ ¼ À rACFðVAcos q cos fÞ 2cos 3 sin3 3 ' $ ( À 1 ðL=DÞ2 þ1 ) : (78) The 3D lift coefficient from Prandtl lifting line theory is CLa ¼ CL0 þ 2p AR AR þ 2 a: (79) The induced drag coefficient from Prandtl lifting line theory is CD ¼ CD0 þ C2 L pARe : (80) From Eq. (79) dCL da ¼ 2p AR AR þ 2 : (81) Combining Eqs. (79) and (80) gives L=D ¼ CL0 þ 2p À AR AR þ 2 Á a CD0 þ À CL0 þ 2p À AR AR þ 2 Á a Á2 pARe (82) Differentiating Eq. (82) gives Table 4 Effect of aspect ratio on aerodynamic force. AR’ 3 [ ] L/D FA [104 N] d(FA)/d(AR) [104 N] 1 19.86 2.77 4.34 5.25 2 13.37 4.21 9.36 4.60 3 11.11 5.09 13.46 3.62 4 9.97 5.69 16.67 2.84 5 9.29 6.12 19.21 2.26 6 8.83 6.44 21.24 1.83 7 8.50 6.69 22.91 1.51 8 8.25 6.90 24.29 1.27 9 8.06 7.06 25.46 1.08 10 7.90 7.20 26.46 0.92 dðL=DÞ da ¼ 0 @CD0 þ À CL0 þ 2p À AR AR þ 2 Á a Á2 pARe 1 A$2p AR AR þ 2 À CL0 þ 2p AR AR þ 2 a $ 0 @ 2CL0 $2p À AR AR þ 2 Á pARe þ 8p2 À AR AR þ 2 Á2 a pARe 1 A 0 @CD0 þ À CL0 þ 2p À AR AR þ 2 Á a Á2 pARe 1 A 2 (83) Fig. 12. Trajectories shapes with different pole circle sizes. Fig. 13. Trajectory shapes with different pole circle separation. G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e26782676
  • 11. Using AR ¼ 4.86, CD0 ¼ 0:05, CL0 ¼ 0:337, ra ¼ 1.19, A ¼ 300, V ¼ 6:18, q ¼ 15, f ¼ 0, and e ¼ 0.9 Eq. (76) was found yield a ¼ 15.28. This is the theoretical optimum as determined by Prandtl lifting line theory, though there are other factors which might influence the best angle of attack for practical use. In particular, the way that the kite recovers from off-design flight conditions, such as stall, is likely to be significantly improved by mounting the kite to operate at a smaller angle of attack. 6.5. Pole separation and circle size The effects of maneuver pole separation and pole circle size were investigated by comparing a series of different kite trajecto- ries (See Figs. 12 and 13). Results were obtained through imple- mentation of the flow chart in Fig. 7, including the effect of the wind gradient as determined by Eq. (46). The effect of increasing pole circle size and pole circle separation on time averaged force output is shown in Figs. 14 and 15 respectively. These figures show that increasing pole circle size and separation both lead to a marked reduction of time averaged propulsive drive force since the maneuver is moved away from its optimal position for maximized propulsive effect. The limitation of these investigations is that no account is taken of the potential reduction in lift coefficient asso- ciated with canopy deformation and angle of roll during a tight turn. 7. Experimental validation Previously Eq. (45), for obtaining onset velocity as a function of kite position, was validated by comparing predicted aerodynamic force with measured kite line tension [7]. The kite forces, force direction and wind speed had been recorded by flying a kite from a dynamometer in natural winds. In the present paper, the implementation of the theory has been developed to allow the kite velocity and motions to be computationally predicted, as well as the instantaneous force based on position. For experimental validation of the present work, two trajectories are shown in Fig. 16. The theoretically parameterized trajectory is defined according to Section 3, whilst the experimentally determined trajectory was recorded using a purpose built kite dynamometer described by Dadd [7]. The theoretically determined shape was selected to closely match that of the experimental one (see Fig. 16). Small differences are apparent since the manually flown trajectory exhibits irregularities. The aerodynamic force was determined in three ways (see Fig. 17). First, the zero mass theoretical aero- dynamic force was determined using the experimentally deter- mined position coordinates as an input. Second, the actual aerodynamic kite force was obtained using the dynamometer. Comparing these two results serves to validate Eq. (45), which determines kite onset velocity as a function of position, as described by Dadd et al [7]. The third result in Fig. 16 is also based on Eq. (45), but uses Eqs. (51)e(59) to establish the kite velocity and hence to theoretically determine the results to a base of time, by integrating around the maneuver. It can be seen that the theoretical time for the kite to progress around the maneuver closely matches that of the experimental kite trajectory. This serves to validate Eqs. (51)e(59). The errors arising through determination of kite force through each of these methods is Fig. 14. Force reduction with circle size. Fig. 15. Force reduction with pole separation. Fig. 16. Experimentally and theoretically defined trajectories for comparison. Fig. 17. Comparison between theoretical predictions and experimental measurements. G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e2678 2677
  • 12. discussed in detail by Dadd et al [7]. The most significant error for experimentally determined kite trajectories was shown to be due to spatial and temporal differences in wind speed across the flight envelope which was not properly accounted for. 8. Conclusions This paper has presented a scheme for predicting time averaged kite forces together with a method for prameterizing figure of eight shape kite trajectories. The theoretical model was modified to allow a high degree of control over input maneuver paths in order to assess the effects of maneuver orientation, shape and true wind angle. The influence of the changing wind strength in the Earth’s natural boundary layer is included. The scheme has been imple- mented to show variation of onset velocity, kite velocity, onset angles and aerodynamic force for an example maneuver. In a case study with a 300 m2 kite, 300 m long flying lines and 6.18 msÀ1 wind speed (at 10 m), peak drive force up to 30 Â 104 N was pre- dicted. A typical vertical figure of eight maneuver in the same conditions was shown to produce a time averaged horizontal drive force component 16.7 Â 104 N. A complete force polar showing the horizontal component of aerodynamic force for different true wind angles has thus been obtained. The calculation of kite velocity and the propagation of the kite around a theoretically defined maneuver have been validated by comparing a real recorded trajectory to a simulated one. The time scales for a theoretical and practical trajectory are shown to agree favourably, as do the magnitudes of minimum and maximum kite force. The increase in wind with altitude is found to double the aero- dynamic force compared to a hypotheticalcase with nowind gradient. Through considering the wind gradient in combination with kite motions, the optimum mean kite elevation angle for ship pro- pulsion during dynamic flight was found to be 15. Horizontally orientated trajectories are shown to improve downwind perfor- mance by a factor of 1.4 compared to vertical ones, since the maneuver can remain closer to the optimum elevation angle for propulsion. Vertical maneuvers are, however, preferred when maintaining close winded performance is necessary. By combining Prandtl lifting line theory with the zero mass model, the derivative dFA=dðARÞ was calculated at a variety of different aspect ratios. For the downwind dynamic flight condition, increasing the kite aspect ratio from 4 to 5, for example, was found to yield a 15% increase in drive force showing that force is highly sensitive to aspect ratio through its effect on kite L/D. Low aspect ratios may however be favourable to kite stability characteristics. The angle of attack at which the kite operates modifies the aerodynamic force generated through its influence on lift and drag coefficients, but also through its effect on L/D. Optimizing the angle of attack with respect to both of these yields a ¼ 15. Optimal trajectories were found to be those that are centred most closely to the best region of the flight envelope. Increasing maneuver pole circle size and pole separation both diminished the time averaged propulsive drive force arising. If the potential effect of reduced CL during tight turns is negligible, the best maneuver is found to be a figure of eight shape with small circular ends and small pole separation, centred closely to an elevation of 15. References [1] Loyd ML. Crosswind kite power (for large-scale wind power production). AIAA Journal of Energy 1980;4(3):106e11. [2] Wellicome, JF, Wilkinson S. Propulsive kites e an initial study. University of Southampton, Department of Ship Science, Faculty of Engineering and Applied Science. Tech. Rept. SSSU 19; 1984. [3] Naaijen P, Koster V, Dallinga RP. On the power savings by an auxiliary kite propulsion system. International Shipbuilding Progress 2006;53(4):255e79. [4] Williams P. Optimal wind power extraction with a tethered kite. AIAA Guidance, Navigation, and Control Conference and Exhibit. AIAA, Keystone, Colorado; 2006. [5] Argatov I, Rautakorpi P, Silvennoinen R. Estimation of the mechanical energy output of the kite wind generator. Renewable Energy 2009;34(6):1525e32. [6] Argatov I, Silvennoinen R. Energy conversion efficiency of the pumping kite wind generator. Renewable Energy 2010;35(5):1052e60. [7] Dadd GM, Hudson DA, Shenoi RA. Comparison of two kite force models with experiment. Journal of Aircraft 2010;47(1):212e24. [8] Claughton AR, Shenoi RA, Wellicome JF. Sailing yacht design theory. 1 ed. Harlow: Addison Wesley Longman; 1998. pp. 5e6. [9] Matlab. Mathmatics laboratory, software package [Ver. 7.0.0]. Cambridge: MathWorks; 2004. [10] Abbott IH, Doenhoff AE. Theory of wing sections. New York: Dover Publica- tions; 1959. [11] Bryant, LW, Brown WS. Collected researches on the stability of kites and towed gliders. National Phisics Laboratory, Reports and Memoranda 2303; 1942. G.M. Dadd et al. / Renewable Energy 36 (2011) 2667e26782678