You are on page 1of 11

Performance Characteristics of a

1 kW Scale Kite-Powered System


A 1 kW scale kite-powered system that uses kites to convert wind energy into electrical
David J. Olinger energy has been studied to determine its performance characteristics and establish fea-
Department of Mechanical Engineering, sibility of steady-state operation. In this kite-powered system, a kite is connected to a
Worcester Polytechnic Institute, tether that transmits the generated aerodynamic forces on the kite to a power conversion
Worcester, MA 01609 system on the ground. The ground-based power conversion system consists of a rocking
e-mail: olinger@wpi.edu arm coupled to a Sprag clutch, flywheel, and electrical generator. Governing equations
describing the dynamical motion of the kite, tether, and power conversion mechanism
Jitendra. S. Goela were developed assuming an inflexible, straight-line tether. A steady-state analysis of the
Optical and Ceramic Technologies, kite aerodynamics was incorporated into the dynamical equations of the kite-powered
Dow Chemical Company, system. The governing equations were solved numerically using a Runge–Kutta scheme
Marlboro, MA 01752 to assess how performance parameters of the system such as output power, cycle time,
e-mail: jgoela@dow.com and tether tension varied with wind speed, kite area, and aerodynamic characteristics of
the kite. The results showed that a 1 kW scale system is feasible using the proposed
design concept with a kite area of 25 m2 and wind speeds of 6 m/s. Preliminary efforts to
build and test a working 1 kW scale kite-powered demonstrator are also reported.
关DOI: 10.1115/1.4002082兴

Keywords: wind energy, alternative energy, kite-powered system

1 Introduction and tethers should also be less compared with rigid blades. Dis-
Wind turbines 共or windmills兲 have traditionally been used to advantages for kite-powered systems include kite control issues
extract power from the wind. Tethered wind energy systems have and the requirement of some type of buoyancy system to keep the
been given less attention, especially in the past 2 decades, al- kite aloft when wind velocity reduces to a low value or is zero.
though several investigations were conducted in late 1970s, early In this study, we consider a kite-powered system consisting of a
1980s 关1–11兴, and more recently in 2000s 关12–14兴. In a tethered large kite 共area ⬃25 m2兲 as the aerodynamic body, a tether, and a
energy system, an aerodynamic body 共kite, parafoil, aircraft, para- power conversion system located on the ground, as shown in Fig.
chute, or balloon兲 is held aloft by the wind. In this study, a kite is 1. The kite tether is attached to one end of the rocking arm so that
connected to a flexible tether, which is used to transmit generated the up-and-down motion of the kite is converted into an oscillat-
aerodynamic forces to a power conversion system on the ground. ing motion of the arm about the pivot point B. The oscillating
In some proposed schemes, the power conversion also takes place motion of the arm is then converted into rotary motion of a fly-
on the aerodynamic body itself 关1–5兴. wheel through the use of a Sprag clutch mechanism, a retraction
Kite-powered systems have certain potential advantages over spring, and several reduction gears, as shown in Fig. 1共a兲. The
wind turbine systems. A main advantage is that kites can fly at
Sprag clutch 共essentially a rear bike hub mechanism兲 engages the
greater heights than those at which wind turbines operate. Some
flywheel during kite ascent to allow generated power to raise the
investigators have proposed flying kites at altitudes of several
thousand feet 关6,7兴. Wind speed increases with height in Earth’s load weight. The Sprag clutch disengages during kite descent. The
boundary layer, and since wind power density is proportional to flywheel serves to smooth out rotation rate variations. Additional
the cube of wind speed, the available wind power is greater for details of the power transmission system are provided in Sec. 2.
kites of equivalent area. Here, a wind turbine and kite are consid- Many previous studies of wind power systems focused on high
ered to be of equivalent area when the kite planform area equals altitude systems, often operating in the jet stream, capable of
the area swept by the turbine blades. large-scale 共megawatt size兲 power production 关2–5兴. Studies have
Other potential advantages for kite-powered systems include also appeared where kites operate in a yo-yo configuration and the
reduced capital costs, simplicity in design and construction, and energy is generated as the kites unroll the lines from reels during
potential for use in regions with lower wind speeds. These advan- the power phase and electric motors, attached to the lines, winding
tages are particularly attractive in developing nations where off- the lines back on to the reels by spending a small amount of
grid, distributed power systems can reduce the need for large capi- energy during the recovery phase 关12,13兴. In the present work, the
tal outlays associated with centralized power generation. Another goals are more modest with a focus on smaller-scale power sys-
advantage is reduced environmental impact. A kite-powered sys- tems 共kilowatt size兲 that could be used in developing nations or
tem would reduce visual pollution that often results when wind
rural areas of developed nations. The initial studies of kite-
turbines are sited in attractive natural landscapes. Noise pollution
powered systems at this smaller scale were conducted by Goela et
and bird kill issues associated with wind turbines should also be
reduced for kite systems because birds will be less likely to fly al. 关6–11兴. In particular, equations of motion that describe the
into the tethers or kites due to the kite’s high elevation and rela- dynamics of the kite during a power cycle were formulated 关7兴.
tively small area. In addition, injuries to birds from flexible kites The steady-state solution of these equations were obtained for
both the ascent and descent phases of kite motion by keeping the
primary system variables such as kite velocity and tether inclina-
Contributed by the Solar Energy Division of ASME for publication in the JOUR- tion angles constant. Analytical expressions for power output,
NAL OF SOLAR ENERGY ENGINEERING. Manuscript received February 6, 2009; final
manuscript received April 27, 2010; published online October 4, 2010. Assoc. Editor: tether tension, and inclination angle were developed, and condi-
Spyros Voutsinas. tions for maximum power output were identified for particular

Journal of Solar Energy Engineering Copyright © 2010 by ASME NOVEMBER 2010, Vol. 132 / 041009-1

Downloaded 03 Nov 2010 to 128.205.114.91. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 1 The kite-powered system. „a… The system in the power configuration
when the Sprag clutch and flywheel are engaged. In the idle configuration,
the Sprag clutch and flywheel are disengaged, and the system splits into the
two subsystems shown in „b… and „c…. The structure C in „a… consists of a
constant radius „RC… arc that ensures that the tension force in the chain
connecting gear G to point C always acts perpendicular to the rocking arm
at point C.

cases. In addition, maximum power output was obtained when the focused toward electrical power generation instead of a kite-
kite moves in a cross-wind direction 共perpendicular to the wind兲 powered pump designed to lift loads of water. Furthermore, the
confirming Loyd’s 关1兴 results. present power system uses a fixed length tether versus a variable
Goela et al. 关7–9兴 also performed preliminary wind tunnel stud- tether length in Refs. 关10,11兴. In addition, a steady-state aerody-
ies of several kite models and theoretically studied wind loading namic analysis of the kite lift and drag forces has been used and
effects on kite tethers. Matos et al. 关15兴 also performed wind incorporated in the equations of motion of the kite. Other inves-
tunnel measurements on parafoil geometries and measured lift and tigators 关12–14兴 also recently incorporated aerodynamic analyses
drag coefficients over a range of wind speeds and angles of attack. into studies of kite-powered systems. Goela et al. 关10,11兴 assumed
While these investigations generally focused on the steady-state a constant lift-to-drag ratio L / D = 6 during the ascent phase and an
operation of kite-powered systems, Goela et al. 关10,11兴 also de-
L / D = 2 during the descent phase of the kite. However, as a teth-
veloped and numerically solved the governing differential equa-
tions of motion for a kite-powered pump. Performance parameters ered kite moves 共in ascent or descent兲 with varying velocity, the
such as output power, water-lifting capacity, and kite oscillation orientation of the relative local velocity vector, effective kite angle
period as a function of kite elevation, kite weight, tether angle of of attack, and resultant lift and drag forces 共Fig. 2兲 vary dynami-
inclination, and kite lift-to-drag ratio were determined. cally with time. This dynamic behavior can have a dramatic effect
In this study, the theoretical maximum performance character- on the kite forces, its motion, and system performance parameters
istics of a 1 kW scale kite-powered system appropriate for use in such as power output and tether tension, as shown in Sec. 4.
a developing nation are determined since losses and inefficiencies In the general form, the dynamics of the kite-powered system is
in the tether and power transmission system have been neglected. complex with the kite moving in three dimensions, and the tether
These objectives are similar to the study of Goela et al. 关10,11兴. profile changing continuously in a nonequilibrium condition as the
However, the present kite-powered system is quite different and is kite moves. To accurately model the tether profile, it would be

041009-2 / Vol. 132, NOVEMBER 2010 Transactions of the ASME

Downloaded 03 Nov 2010 to 128.205.114.91. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
actual kite-powered system has four tethers, which will prevent
kite pitching.
This paper is organized as follows. Section 2 describes in detail
the operation of the kite-powered system. In Sec. 3, the governing
equations describing the dynamical motion of the kite, tether,
rocking arm, and power transmission system are developed as-
suming an inflexible, straight-line tether. The numerical scheme
used to solve the dynamics equations is discussed. In Sec. 4, the
solutions of the governing equations are provided to assess how
performance parameters of the system such as output power, cycle
time, and tether tension vary with wind speed, kite area, and aero-
dynamic characteristics of the kite. Finally, a summary of key
results and a working kite-powered demonstrator are briefly de-
Fig. 2 Parameter definitions for the kite and tether scribed in Sec. 5.

required to formulate and solve nonequilibrium wave propagation


equations. This description is beyond the scope of this proof-of- 2 Kite-Powered System
concept study as the objective here is to perform a simplified
analysis aimed at obtaining a steady-state operation that will de- The kite-powered system consists of a kite, a tether, a rocking
velop understanding and guide in producing a working model of arm, and gear system, which converts the wind energy to rota-
the kite-powered system. Consequently, the following simplifying tional energy of the flywheel. The flywheel then powers an elec-
assumptions are made: 共1兲 The analysis of kite energy system is trical generator and converts mechanical energy to electrical en-
based on steady-state aerodynamics, which means that dynamic ergy. To calculate power produced by this system, the electrical
inflow effects are neglected. Dynamic inflow here refers to local generator has been replaced with a load whose potential energy
changes in the flow pattern, the kite effective angle of attack, and increases as the flywheel rotates in a counterclockwise direction
the resultant lift and drag forces due to unsteady aerodynamic 共Fig. 1共a兲兲. A balanced rocking arm with identical lengths on each
effects in the flow. 共2兲 The kite motion is restricted to the two- side of a central pivot, B is used. A counterweight, WCTR, is
dimensional x-z plane 共Fig. 2兲. This assumption will still be able mounted on the rocking arm at point A. A chain attached to a
to provide cross-wind kite motions, as described in Refs. spring is wound around the gear G and then its end is attached to
关1,10,11兴. Disturbances 共wind gusts兲 that could lead to instabilities the rocking arm at point C. The structure C in Fig. 1共a兲 consists of
in kite motion are also not considered here. 共3兲 The kite tether is a constant radius 共RC兲 arc that ensures that the tension in the chain
inflexible and its profile is a straight-line. This assumption is valid connecting gear G to point C will always act perpendicular to the
when wtLt / Ft Ⰶ 1, where wt is the weight per unit length of the rocking arm at point C. The spring and counterweight, WCTR, are
tether, and if aerodynamic forces on the kite tether are small com- used to move the rocking arm down during the kite descent phase.
pared with tether tension. Current kite lines weigh approximately A Sprag clutch 共essentially a rear bike hub mechanism兲 is located
0.015 N/m, and Lt ⬇ 375 m, Ft ⬇ 100 N in our study, so that between gears G and G2, which share a common axle. The Sprag
wtLt / Ft ⬇ 0.058, which is small. The tether transmission effi- clutch is used to disengage 共or engage兲 gear G from gear G2,
ciency, eT, is set equal to 1, implying that the tether tension at the which is coupled to the flywheel through gear FG and a drive
two ends 共kite and rocking arm, point A兲 are equal. 共4兲 The kite chain 共Fig. 1共c兲兲. The rocking arm and gear power transmission
tether’s weight, lift, and drag forces are neglected as they are system have two main configurations: 共1兲 the power configuration
small compared with corresponding parameters for kite. For a in which the Sprag clutch and flywheel are engaged and the rock-
tether of diameter 2 mm and wind speed of about 6 m/s, the ing arm moves up 共Fig. 1共a兲兲. In this case, the rocking arm does
aerodynamic forces on the kite tether are small compared with the work against counterweight, stretches the spring, stores energy in
kite lift, drag, and tether tension 关10,11兴. 共5兲 The weight and mo- the flywheel, and increases the potential energy of the load. 共2兲
ment of inertia of the gears, Sprag clutch, connecting chain, and The idle configuration in which the Sprag clutch and flywheel are
drive chain in Fig. 1 are small compared with the weight and disengaged and the rocking arm moves down 共Figs. 1共b兲 and
moment of inertia, respectively, of the flywheel and are also ne- 1共c兲兲. In this case, the counterweight and spring do work on the
glected. 共6兲 A single kite tether is assumed. In a working kite- rocking arm to move it down and the flywheel uses its stored
powered system, additional kite tethers may be needed to control energy to increase the potential energy of the load.
the kite motion and change its angle of attack. 共7兲 The effect of To create the up-and-down motion of the kite, the geometric
dynamics of components that would be used to vary the kite’s angle of attack ␣ of the kite with respect to the horizontal is varied
angle of attack are considered to be small and are not modeled in between two distinct values: ␣ = ␣asc and ␣ = ␣dsc with ␣asc
this study. 共8兲 No pitching of the kite is allowed except at two ⬎ ␣dsc 共Fig. 2 and Table 1兲. The change in geometric angle of
specific times 共times 3 and 6 in Table 1兲 during a kite cycle. attack is assumed to occur linearly over a specified time period
Although the present analysis is based on one kite tether, the ⌬tAOA Ⰶ TC. In order to keep the analysis simple, the rotational

Table 1 Sequence of events during system cycle

Rocking arm position


Time or movement 共point A兲 Sprag clutch System configuration Angle-of-attack

1 Bottom Disengaged Idle ␣ = ␣asc


2 Ascending Engages Flywheel Power ␣ = ␣asc
3 Ascending Engaged Power Shifts to ␣ = ␣dsc
4 Ascending Disengages Flywheel Idle ␣ = ␣dsc
5 Top Disengaged Idle ␣ = ␣dsc
6 Descending Disengaged Idle Shifts to ␣ = ␣dsc

Journal of Solar Energy Engineering NOVEMBER 2010, Vol. 132 / 041009-3

Downloaded 03 Nov 2010 to 128.205.114.91. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
At time 3, where the rocking arm angle ␥ = ␥dsc, the angle-of-
attack mechanism shifts the geometric angle of attack of the kite
to ␣ = ␣dsc to reduce the kite lift. The lower ␣dsc value decreases
the lift on the kite, and slows the upward motion of the arm. At
time 4 the Sprag clutch disengages from the flywheel 共␻G
⬍ ␻FRFG / RG2兲 and the flywheel’s angular velocity decreases 共Fig.
3共b兲兲. At time 5, point A on the rocking arm reaches its maximum
height and reverses direction due to the effect of the counter-
weight WCTR and the retraction spring. The gear G now rotates in
a clockwise direction 共negative angular velocity values in Fig.
3共b兲兲 as the rocking arm moves downwards. As the rocking arm
共point A兲 descends, the retraction spring maintains the tension in
the chain connecting the rocking arm to gear G. In addition, the
stored energy in flywheel continues to rotate gears G2 and FG in
a counterclockwise direction and move the load weight upward. It
is to be noted that in this study the kite dynamics and rocking arm
motion are fully modeled as the rocking arm undergoes transition
from an ascending motion to a descending motion at time 5. The
maximum rocking arm angle 共approximately ␥ = 40 deg兲 is not
specified as a simulation input but is determined automatically
from the solutions of the equations of motion.
At time 6, where the rocking arm angle ␥ = ␥asc, the angle-of-
attack mechanism shifts the kite’s angle-of-attack to ␣ = ␣asc.
However, the momentum of the rocking arm continues to move
the arm down until a minimum rocking arm angle is obtained at
time 1. This completes the system cycle and the system starts all
over again at time 1.

3 Governing Equations
In this section, the governing differential equations of motion
for the kite, rocking arm, and flywheel are developed. The primary
system variables are defined in Figs. 1 and 2. Simplifying assump-
tions were discussed in Sec. 1.
Figure 2 summarizes the forces and moments acting on the kite
that result from the kite motion. The kite is restricted to two-
dimensional motion in the x-z plane. The kite moves with a ve-
locity VK at angle ␤ with respect to the ground. The relative or
local velocity VR at an angle of ␾ with respect to the ground
Fig. 3 „a… Rocking arm motion versus time. „b… Angular veloci- results from a combination of the kite motion and the horizontal
ties of the rocking arm, gear G, and the flywheel versus time. wind velocity, as shown in the velocity triangle in Fig. 2. The kite
tether is oriented at angle ␪ with respect to the ground. The lift
and drag forces FLK and FDK are oriented perpendicular and par-
allel to the VR vector, respectively. The force on the tether Ft is
dynamics of the kite, which would require consideration of the considered constant along the tether length since tether transmis-
kite pitching moment about the aerodynamic center, are not mod- sion losses due to tether profile are neglected.
eled. The wind velocity variation with height above the ground is
Table 1, Fig. 1, and Fig. 3 provide the sequence of events that specified by setting a reference wind velocity V0 at a reference
occur during a full cycle of the system. At time 1, point A on the height of z0 = 10 m and using
end of the rocking arm is at its minimum height, and the Sprag
clutch and flywheel are not engaged 共idle configuration兲. At time
1, ␣ = ␣asc yields a higher lift force on the kite, which accelerates
V = V0 冉 冊
zkite
z0
1/7
共2兲

the rocking arm upward and rotates gear G in a counterclockwise from Ref. 关16兴 where zkite = Lt sin ␪ is the elevation of the kite.
direction. The angular velocity of the rocking arm and gear G The velocities V1 and V1A represent the kite velocities along the
increases until the Sprag clutch engages with the flywheel 共power kite tether direction with respect to the stationary ground and
configuration兲 at time 2 共Fig. 1共a兲兲. This engagement occurs when point A 共end of rocking arm兲 respectively. Likewise, the velocities
V2 and V2A represent the kite velocities normal to the kite tether
RC RFG
␻G = ␻ ⱖ ␻F 共1兲 with respect to the stationary ground and point A, respectively.
RG RG2
3.1 Kite Dynamics. Governing equations for the kite dynam-
One can clearly see the engagement of the Sprag clutch and fly- ics are derived as follows. Taking moments of the kite drag, lift,
wheel at time 2 in Fig. 3共b兲, which shows the angular velocity as and weight about point A at the end of the rocking arm, and
function of time of rocking arm, gear G, and flywheel during a dividing by the tether length yield
single cycle of the rocking arm. The rate of change of the angular
velocity ␻ decreases suddenly since the rocking arm must now WK dV2A 2WKV1AV2A
− = FDK sin共␪ + ␾兲 − FLK cos共␪ + ␾兲
overcome additional inertia of the flywheel and load weight. g dt gLt
When the Sprag clutch and flywheel are engaged, gears G, G2,
FG, and the flywheel all rotate in a counterclockwise direction + WK cos ␪ 共3兲
共Fig. 1共a兲兲 and the load weight moves upward. And taking a force balance along the tether yields

041009-4 / Vol. 132, NOVEMBER 2010 Transactions of the ASME

Downloaded 03 Nov 2010 to 128.205.114.91. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
2
d␻F
兺M
WK dV1A WKV2A F CR G
= FDK cos共␪ + ␾兲 + FLK sin共␪ + ␾兲 − WK sin ␪ + IF = F = FFGRFG − WLOADRF = RFG − WLOADRF
g dt gLt dt RG2
− Ft 共4兲 共11兲
The above two equations apply when the system is in the power
In the above equation, the tether’s moment of inertia, lift, and drag
configuration. The following equations are used in Eq. 共11兲:
are neglected as discussed in Sec. 1 earlier. In Eq. 共3兲, on the
left-hand side, the respective terms are due to the inertial and RFG
Coriolis acceleration of the kite and the tether, while on the right- RC␻ = RG␻G = RG␻G2 = RG ␻F 共12兲
RG2
hand side the respective terms are due to the kite drag, lift, and
weight. In Eq. 共4兲 from left, we have terms due to kite inertia, FCRG␻G = FG2RG2␻G2 = FFGRFG␻F 共13兲
drag, lift, weight, and centrifugal acceleration, and the tether ten-
sion. The Coriolis term in Eq. 共3兲 and inertia term in Eq. 共4兲 drop Here, FFG = FG2 is the tension in the “tight” drive chain between
out since, in the present case, the tether length Lt is constant so gears G2 and FG, and FC is the tension in the chain from the
that V1A = 0. rocking arm to gear G. The tension in the “loose” following chain
The kite lift and drag forces FLK, FDK are calculated assuming between gears G2 and FG is assumed to be negligible.
that the kite is modeled as a finite length wing 共in the spanwise By eliminating the tension force FC between the two moment
direction兲 with a thin airfoil section. Assuming a linear lift coef- balances in Eqs. 共10兲 and 共11兲, and using Eqs. 共12兲 and 共13兲 and
ficient curve 共below airfoil stall angle兲, the lift coefficient is given VA = RA␻, an equation for the linear acceleration of point A is
by 关17,18兴 obtained. The moments of inertia of the rocking arm around the
pivot B, and the flywheel are given by
a0
CL = 共␣ − ␾ − ␣L=0兲 共5兲 1 WDA共2RA兲2 WCTRRCTR
2
a0 IAD = + 共14兲
1+ 12 g g
␲共AR兲
1 WFRF2
where ␣L=0 is the angle of attack at which zero lift occurs on the IF = 共15兲
kite. 2 g
The kite drag coefficient is calculated using The Sprag clutch in the power transmission mechanism can either
be engaged or disengaged 共see Table 1兲. It is disengaged 共idle
CL2 configuration兲 when ␻G ⬍ 共RFG / RG2兲␻F in which case IF = Wload
CD = CD0 + 共6兲
␲共AR兲e0 = 0 in Eqs. 共10兲 and 共11兲. While disengaged the flywheel angular
velocity is determined through
from finite wing theory 关17,18兴 where CD0 is a parasitic drag
coefficient that introduces viscous effects into the drag calcula- d␻F
IF = − WLOADRF 共16兲
tion, and e0 is an Oswald efficiency factor that accounts for drag- dt
due-to-lift effects and wing tip vortex effects from nonelliptic
When the Sprag clutch is engaged 共power configuration兲, ␻G
shaped wings. The lift-to-drag ratio L / D, which also varies with
ⱖ ␻FRFG / RG2, and IF, Wload are set to their specified input values
time during the kite motion, is given by CL / CD. The pitching
in Eq. 共10兲 and 共11兲, and ␻F = ␻GRG2 / RFG.
moment about the kite aerodynamic center is not included in the
In addition, the following kinematic equation is required:
model since this moment cannot be transferred through the tether
to the system on the ground. The pitching moment would need to d␥ − VA
be included if the rotational dynamics of the kite between angles = 共17兲
dt RA
␣ = ␣asc and ␣ = ␣dsc were modeled.
The time-varying lift and drag on the kite are then given by Equations 共3兲, 共9兲, 共10兲, 共16兲, and 共17兲 are the five first-order
equations describing the motion of the kite, tether, rocking arm,
1
FLK = 2 ␳VR2 CLAK 共7兲 and power transmission gear system.
The initial conditions for the five first-order differential equa-
and tions at t = 0 are VA = 0, V2A = 0, ␥ = −25 deg, ␪ = 70 deg, and ␻F
= 0. The differential equations are solved for VA, V2A, ␥, ␪, and ␻F
1
FDK = 2 ␳VR2 CDAK 共8兲 as functions of time using a Runge–Kutta scheme in a MATLAB
algorithm. Studies were conducted to ensure that the numerical
In addition, we require the differential equation results were independent of initial conditions and numerical time
step ⌬t. Once these five primary variables are determined at each
d␪ − V2A time step 共with ⌬t = 0.002 s兲 the following equations are used in
= 共9兲 order to calculate other important system parameters:
dt Lt
V1 = VA sin共␪ + ␥兲 共18兲
3.2 Rocking Arm and Power Mechanism Dynamics. To
model the motion of the rocking arm, a moment balance about the V2 = VA cos共␪ − ␥兲 + V2A 共19兲

冋册
pivot point B in Fig. 1 is applied
V2
␤ = ␪ − tan−1 共20兲
d␻
IAD
dt
= 兺 M B = FtRA cos共␥ − ␪ + ␲/2兲 − FCRC − K⌬xRC
V1

VK = 冑V21 + V22 共21兲


RA RD
− WBA cos共␥兲 + WDB cos共␥兲 − WCTRRCTR cos共␥兲
2 2 VR = 冑V2 + VK2 + 2VVK cos ␤ 共22兲
共10兲
and a second moment balance about the flywheel center 共point F兲. ␾ = tan−1 冋 VK sin ␤
V − VK cos ␤
册 共23兲

Journal of Solar Energy Engineering NOVEMBER 2010, Vol. 132 / 041009-5

Downloaded 03 Nov 2010 to 128.205.114.91. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
3.3 Power Output. The average power output of the system 4 Results
is determined using
In Table 2, the input parameters for a baseline run of the nu-
WLOAD共⌬zLOAD 2−4 + ⌬zLOAD 4−2兲
merical simulation are presented. An AK = 25 m2 kite with an as-
P̄K = 共24兲 pect ratio of AR = 4 and a cambered airfoil 共␣L=0 = −4 deg兲, flying
TC
at an elevation of Lt = 375 m, with a nominal 共design兲 wind speed
where ⌬zLOAD 2−4 = 兺TT4RF␻F⌬t and ⌬zLOAD 4−2 = 兺TT2RF␻F⌬t are of V0 = 6 m / s at an elevation of z0 = 10 m is studied. The load and
2 4
the elevation changes of the load weight during the engaged and counterweight are set to WLOAD = 80 N and WCTR = 300 N, re-
disengaged segments 共Table 1兲 of the system cycle, and TC is the spectively. Kites of AK = 25 m2 and aspect ratio AR = 4 are com-
time period for 1 cycle of the system. The power output is aver- mercially available. Also, several smaller kites could be “stacked”
aged over approximately 15 system cycles at times after initial above each other in series on a single set of tether lines to achieve
transients have died out. the required kite area. Kites of this size can be flown at an eleva-
Energy balances were conducted over a full cycle of the system tion of 375 m. Startup issues such as initial uncoiling of the tether
to validate the power output results. This study confirmed that lines and the elevation of the kite to the required height will be
each of the following average power calculations 共for the kite, considered in future publications. A Class 4 wind of speed V0
pivot B, and gear G in Eqs. 共25兲–共27兲, respectively兲 matched the = 6 m / s has been modeled here. Gear mechanisms, their radii,
average power output given in Eq. 共24兲.
TC

兺FV
1
P̄K = t K cos共␪ − ␤兲⌬t 共25兲
TC 0

TC

兺 F R ␻ cos共␥ − ␪ + ␲/2兲⌬t
1
P̄B = t A 共26兲
TC 0

TC

兺F R
1
P̄G = C G␻G⌬t 共27兲
TC 0

In addition, the instantaneous power can be determined by calcu-


lating
PK共t兲 = FtVK cos共␪ − ␤兲 共28兲

PLOAD共t兲 = WLOADRF␻F 共29兲


for the kite or load.
A nondimensional power coefficient is given by


CP = 共30兲
0.5␳V̄3AK
Power coefficients will be correlated against the normalized load
weight given by

ⴱ WLOAD
WLOAD = 共31兲
0.5␳V̄2AK

Table 2 Input parameters for baseline run

AK = 25 m2 RG = 0.05 m
AR = 4 RG2 = 0.05 m
a0 = 2␲ 1/rad V0 = 6 m / s
b = 10. m WBA = 125 N
CD0 = 0.1 WCTR = 300 N
c = 2.5 m WDB = 125 N
eo = 0.9 WF = 462 N
g = 9.81 m / s2 WK = 55 N
IAD = 244.2 kg m2 WLOAD = 80 N
IF = 0.924 kg m2 z0 = 10 m
K = 10 N / m ␣asc = 20 deg
Lt = 375 m ␣dsc = −20 deg
RA = 2.5 m ␣L=0 = −4 deg
RC = 2.5 m ⌬t = 0.002 s
RCTR = 2.5 m ⌬tAOA = 0.25 s
Re= 1.0⫻ 106 ␥asc = −25 deg
RF = 0.2 m ␥dsc = 25 deg
RFG = 0.025 m ␳ = 1.21 kg/ m3 Fig. 4 Time variation of rocking arm and power mechanism
parameters. „a… ␥, ␻F, and ␻. „b… Ft. „c… PK„t… and PLOAD„t….

041009-6 / Vol. 132, NOVEMBER 2010 Transactions of the ASME

Downloaded 03 Nov 2010 to 128.205.114.91. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
and various input weights are selected based on their correspond- per kite tether, which is well below the tensile strength of many
ing parts in the demonstrator system described in Sec. 5. available kite tethers. However, the sudden changes in tether ten-
sion imply that a fatigue analysis on the tethers and structure will
4.1 System Dynamics. Figure 4 presents the time variation of need to be considered as part of future system design.
several rocking arm and power mechanism parameters. In Fig. Instantaneous power produced by the kite 共Eq. 共25兲兲 and load
4共a兲 the stable, periodic oscillation of the rocking arm angle ␥ is weight 共Eq. 共24兲兲 are presented in Fig. 4共c兲. The kite curve shows
observed with a period of approximately 3.0 s. The system a positive net power output since the positive power produced
reaches the stable oscillation approximately 15 s after the kite is during the kite ascent is greater than the negative power output
set in motion. The angular velocity of the arm ␻ is about 1.0 rad/s during descent. The load weight power curve shows how the fly-
as the arm ascends and the flywheel is engaged. The flywheel wheel stores and distributes energy to maintain a positive power
angular velocity averages approximately 65 rad/s. output throughout the cycle. Figure 4共c兲 also shows that the de-
In Fig. 4共b兲, the variation of kite tether tension with time is sired 1 kW power level is obtainable with the kite-powered sys-
presented. Tether tensions of approximately 2000 N and 100 N are tem under study.
observed during the ascent and descent phases of the arm. Peak Figure 5 presents the time variation of several kite parameters.
tether tensions reach approximately 2500 N at time 6. This analy- The tether angle oscillates around a mean of ␪ ⬵ 70 deg. Goela
sis assumes a single kite tether; however, in actual kite-powered 关7兴 showed that the tether profile can be considered as a rigid,
systems, up to four tethers may be used in order to control the kite straight line for tether inclination angles of 70 deg and tether
motion. This reduces peak tether tensions to approximately 500 N lengths of Lt = 375 m, justifying the assumption in this study. The
sudden changes in the kite motion angle ␤ in Fig. 5共a兲 are con-
sistent with the kite motion in the x-z plane, as shown in Fig. 5共c兲.
This figure presents kite locations 共relative to the rotating arm
axis, point B兲 for times after initial transients have died out. Fig-
ure 5共c兲 clearly shows the highly periodic, stable motion of the
kite during a power cycle. Figure 5共a兲 also presents the time varia-
tion of the effective angle of attack ␣eff of the kite, which varies
between ␣eff ⬵ 10 deg during times 2-4 and ␣eff ⬵ −2 deg during
times 4-2. These values are well within the linear range of lift
coefficient, Eq. 共5兲, for typical airfoils. This validates use of the
linear aerodynamic theory. The rapid transients in kite motion
angle ␤ and effective angle of attack ␣eff of the kite suggest that
dynamic inflow effects need to be included in subsequent, more
detailed analysis. A reduced frequency c / TCV̄ ⬇ 0.08 can be cal-
culated based on c = 2.5 m, TC = 3 s, V̄ = 10 m / s, which is within
the range where dynamic inflow becomes important. As discussed
in Sec. 1, modeling of dynamic inflow effects is beyond the scope
of this proof-of-concept study. These effects may lead to instabili-
ties in the kite motion, which will be studied in future analysis.
The kite lift-to-drag ratio, L / D, is shown in Fig. 5共b兲. This ratio
varies between values of 5 and 2 during times 2-4 and 4-2, re-
spectively. These values are close to L / D values assumed by
Goela et al. 关10,11兴. Figure 5共b兲 also presents velocities VK and
VR, which reach peak values of approximately 5 m/s and 15 m/s,
respectively. The corresponding average wind velocity at the kite
elevation for this case is approximately 10 m/s from Eq. 共2兲,
showing that the kite motion significantly increases local relative

Fig. 5 Time variation of kite parameters. „a… ␪, ␤, and ␣eff. „b…


VR, VK, and L / D. „c… Kite position showing stable oscillation.
The kite position is shown for times after initial transients have
died out. The numbered times are defined in Table 1 and Fig. 3. Fig. 6 Power coefficient versus wind velocity

Journal of Solar Energy Engineering NOVEMBER 2010, Vol. 132 / 041009-7

Downloaded 03 Nov 2010 to 128.205.114.91. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
wind velocities and resultant forces on the kite. The Reynolds
number 共based on kite chord length and V0兲 is 1.0⫻ 106 for the
baseline run; however, the local Reynolds number 共based on VR兲
varies with time over a range of 8.0⫻ 105 ⬍ Re⬍ 2.5⫻ 106 for the
V0 = 6 m / s case given the changes in local relative wind velocity
in Fig. 5共b兲.
4.2 Variation in Design Parameters. In this section the
variation in average power output with various system design pa-
rameters is studied. Unless otherwise specified, the system values
for the baseline run of Table 2 were used.
In Fig. 6, the effect of wind velocity on normalized power
output is presented. The shape of the power curve resembles that
of conventional wind turbines. Peak power coefficients occur for
5 ⬍ V0 ⬍ 6 m / s, e.g., from class 2 to class 4 winds.
Figure 7 presents the power coefficients as a function of both

normalized load weight, WLOAD , and average flywheel angular
velocity, ␻
¯ F, for three different wind velocities: V0 = 5, 6, and 7
m/s with tether length held constant. The three velocities studied
constitute class 2, 4, and 6 winds, respectively. Higher wind
speeds produce more power as expected but also produce a higher
value of power coefficient for most load weights selected. The
flywheel angular velocity is inversely dependent on the specified
magnitude of the load weight. For wind velocity, V0 = 6 and 7 m/s,
the power coefficient curves peak with normalized load in the

range of 0.06⬍ WLOAD ⬍ 0.08, and with the average flywheel an-

Fig. 8 Effect of kite area on power output. „a… Power coeffi-


cient versus normalized load weight. „b… Power coefficient ver-
sus average flywheel angular velocity. V̄ = 9.96 m / s.

gular velocity in the range of 50–70 rad/s. These results show that
kite-powered system loads can be optimized to achieve efficient
performance for specified system operating parameters.
Figure 8 shows the variation of normalized power with kite area
and average flywheel angular velocity. For these calculations,
when the kite area was varied, the kite weight was also corre-
spondingly scaled. The power coefficient curves show a small
variation with different size kites, which indicates that the average
power from the kite system will scale with kite area through Eq.
共30兲.
To further understand the variations in power levels with load
weight, a study of cycle, engagement 共2-4兲 and disengagement
共4-2兲 times was conducted. The system values for the baseline run
of Table 2 were used. As shown in Fig. 9, the cycle time for the
system varies by a small amount around TC = 3.0 s as the load
weight varies. However, the time T2−4 during which the Sprag
clutch and flywheel are engaged increases with increasing load

weight. Near the optimum conditions at WLOAD ⬇ 0.06, the duty
cycle is approximately 50% since T2−4 ⬵ T4−2.
Figure 10 shows the effect of variation in the counterweight on
normalized output power. Average power levels generally increase
with decreasing counterweight as expected since less of the gen-
erated power from the kite is used to lift the counterweight at the
rocking arm end. The relatively minor variation in power levels
suggests that the system can operate over a wide range of coun-
terweight values.
Fig. 7 Effect of wind velocity on power output. „a… Power co- In Figs. 11 and 12, the effects of variations in different kite
efficient versus normalized load weight. „b… Power coefficient parameters on power coefficient are presented. Figure 11 shows
versus average flywheel angular velocity. V̄ = 8.31, 9.96, 11.6 that increasing the camber of the kite airfoil increases the power
m/s for V0 = 5.0, 6.0, 7.0 m/s cases, respectively. output quite dramatically as expected. Figure 12 shows the effect

041009-8 / Vol. 132, NOVEMBER 2010 Transactions of the ASME

Downloaded 03 Nov 2010 to 128.205.114.91. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 9 System period TC, engagement time T2−4, and disengagement time T4−2 versus
normalized load for the baseline run of Table 2.

of kite aspect ratio. Higher aspect ratio kites yield higher output The effect of various system variables such as wind velocity,
power due to decreased wing tip vortex effects and induced drag. load weight, and various kite parameters 共area, airfoil camber, and
Relatively small changes in camber and aspect ratio result in sig- aspect ratio兲 on the theoretical maximum power output was deter-
nificant changes in power coefficient levels. These results suggest mined. The results showed that a 1 kW scale system is feasible
that optimization of the kite airfoil and wing planform shape using the proposed design concept with a kite area of 25 m2 and
should benefit system performance. This optimization analysis wind speeds of 6 m/s based on theoretical maximum performance
will be the subject of another study in the future. characteristics. The system inefficiencies pertaining to the tether
and power conversion mechanism were not included in the above
5 Conclusions calculations. If an overall system efficiency of 60% or higher is
assumed, the results showed that 1 kW power can still be obtained
The performance characteristics of a 1 kW scale kite-powered using a kite area of 25 m2 and wind speeds of V0 = 7 m / s since
system appropriate for sustainable development in an impover- the average power levels reached approximately 1800 W for these
ished nation were determined. The kite-powered system consists conditions.
of a large kite, a tether, and a ground-based energy conversion The present study has provided a proof-of-concept analysis of a
mechanism. Governing equations describing the dynamical mo- 1 kW scale kite-powered system. Extension of this work in the
tion of the kite, tether, rocking arm, and power transmission sys- future will include conducting wind tunnel test on scale-models of
tem were developed assuming an inflexible, straight-line tether. A various kite geometries with the aim of obtaining lift and drag
steady-state, aerodynamic analysis using linear aerodynamic data over a range of Reynolds numbers. These data will then be
theory was used to account for the variation in lift and drag coef- incorporated into the dynamical equations of motion of the kite
ficients with angle of attack. The governing equations were solved energy system to study the power variations with Reynolds num-
numerically using a Runge–Kutta scheme. ber. An analysis of instabilities and transient behavior due to sud-
den changes in wind speed is another area for future work.

Fig. 10 Effect of counterweight on power output. Power coef- Fig. 11 Effect of kite airfoil camber on power output. Power
ficient versus normalized load weight. V̄ = 9.96 m / s. coefficient versus normalized load weight. V̄ = 9.96 m / s.

Journal of Solar Energy Engineering NOVEMBER 2010, Vol. 132 / 041009-9

Downloaded 03 Nov 2010 to 128.205.114.91. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
energy captured by the system into direct-current 共dc兲 electricity.
The output from the generator is stored in eight Deka/MK® AGM
storage batteries with 12 V potential and 32.5 A h capacity per
battery. When alternating-current 共ac兲 electricity is needed, a
Xantrex® Trace 3624 inverter converts the dc electricity stored in
the batteries. A Morningstar Tristar® TS-60 charge controller
guards against overcharging of the batteries while monitoring
storage levels. The electrical system is not modeled in the present
study.
The kite-powered system has undergone initial field testing and
generated electrical power for a short period of time. In future
work, it is planned to compare model predictions from the present
study with field measurements from the demonstrator.

Acknowledgment
The authors gratefully acknowledge the work of the following
WPI undergraduate students on various aspects of kite power: G.
Fig. 12 Effect of kite aspect ratio on power output. Power co- Baldwin, P. Bertoli, M. Blouin, Jr., R. Buckley, C. Coschen, M.
efficient versus normalized load weight. V̄ = 9.96 m / s. DeCuir, M. Hurgin, B. Isabella, T. LaLonde, E. Lovejoy, J. Rod-
den, M. Sangemano, N. Simone, and N. Urko. They also thank A.
Ghezzi for teaching us about high performance kites, and D. Per-
Before closing, work performed at Worcester Polytechnic Insti- kins for his support of our project. This work was funded by the
tute 共WPI兲 to design and construct a working kite-powered dem- U.S. Environmental Protection Agency through the Fourth Annual
onstrator 共Fig. 13兲 is briefly discussed below. This system was P3 Award Program 共P3: People, Prosperity, and the Planet兲, the
developed by the 2 authors of this study and 14 other undergradu- Dow Chemical Co., and Heifer International’s Overlook Farm
ate students in design projects over the past few years. The geo- Learning Center.
metrical parameters of the system closely approximate the corre-
sponding parameters for the baseline run 共Table 2兲 in the present Nomenclature
study.
AK ⫽ kite area
The system uses a Peter Lynn Guerilla® kite of 10 m2 size.
AR ⫽ kite aspect ratio
This kite has four tethers. Two are attached to the leading edge of
the kite and transmit the tether tension to the power mechanism on a0 ⫽ 2D lift curve slope for kite= 2␲ 共1/rad兲
the ground. The other two tethers are connected to the trailing b ⫽ wing span of kite
edge of the kite and allow for control of the kite’s angle of attack c ⫽ chord length of kite
in addition to transmitting tether tension. A simple angle-of-attack CD ⫽ total drag coefficient of kite
change mechanism, based on a sliding weight attached to the CD0 ⫽ parasitic drag coefficient of kite
rocking arm, was designed. The sliding weight pulls or releases CL ⫽ lift coefficient of kite
the tethers connected to kite trailing edges to change the kite’s CP ⫽ power coefficient
geometric angle of attack between ␣asc and ␣dsc. A second mecha- eo ⫽ Oswald efficiency factor
nism based on a four-bar linkage controls the roll motion of the eT ⫽ tether tension transmission efficiency
kite. FC ⫽ tension force in chain from rocking arm to
A power conversion mechanism, consisting of an aluminum gear G
rocking arm, Sprag clutch and spring 共adapted from a rowing FDK ⫽ drag force on kite
machine兲, power transmission gears and chains, and a flywheel, is FFG ⫽ tension in tight drive chain on gear FG
housed within a wooden support structure 共Fig. 13兲. The power FG2 ⫽ tension in tight drive chain on gear G2
conversion mechanism is similar to the mechanism modeled in the FLK ⫽ lift force on kite
present study. An electrical system 共Fig. 13兲 was also designed Ft ⫽ tether tension
and constructed. A WindBlue® model 512 permanent magnetic g ⫽ gravitational acceleration
generator 共attached to axle FG in Fig. 1兲 converts the mechanical IAD ⫽ moment of inertia of rocking arm
IF ⫽ moment of inertia of flywheel
K ⫽ retraction spring constant
L/D ⫽ CL / CD = lift-to-drag ratio
Lt ⫽ tether length
P̄ ⫽ average power output
P共t兲 ⫽ instantaneous power output
RA ⫽ half-length of rocking arm
RC ⫽ radius from pivot to chain attachment point
Re ⫽ kite Reynolds number 共based on chord兲
RCTR ⫽ radius of counterweight
RF ⫽ radius of flywheel
RFG ⫽ pitch radius of gear FG
RG ⫽ pitch radius of gear G
RG2 ⫽ pitch radius of gear G2
T2−4 ⫽ flywheel engagement time 共each cycle兲
T4−2 ⫽ flywheel disengagement time 共each cycle兲
Fig. 13 The kite-powered demonstrator. The rocking arm TC ⫽ time period of system oscillation
structure, kite, power conversion system, battery bank, and t ⫽ time 共s兲
electrical system are shown. V ⫽ horizontal wind velocity at elevation zkite

041009-10 / Vol. 132, NOVEMBER 2010 Transactions of the ASME

Downloaded 03 Nov 2010 to 128.205.114.91. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
V̄ ⫽ average wind velocity at kite= 共1 / TC兺T0 CV⌬t兲 ␻ ⫽ angular velocity of rotating arm
␻F ⫽ angular velocity of flywheel
VA ⫽ velocity of end of rotating arm 共point A兲
VK ⫽ kite velocity with respect to the ground ␻
¯ F ⫽ average angular velocity of
VR ⫽ local relative wind velocity with respect to flywheel= 共1 / TC兲兺T0 C␻F⌬t
ground ␻FG ⫽ angular velocity of gear FG
V0 ⫽ reference horizontal wind velocity at elevation ␻G ⫽ angular velocity of gear G
z0 ␻G2 ⫽ angular velocity of gear G2
V1 ⫽ velocity of kite along tether with respect to
ground
V1A ⫽ velocity of kite along tether with respect to
point A References
V2 ⫽ velocity of kite normal to tether with respect to 关1兴 Loyd, M. L., 1980, “Crosswind Kite Power,” J. Energy, 4共3兲, pp. 106–111.
ground 关2兴 Fletcher, C. A. J., and Roberts, B. W., 1979, “Electricity Generation from Jet
V2A ⫽ velocity of kite normal to tether with respect to Stream Winds,” J. Energy, 3, pp. 241–249.
point A 关3兴 Fletcher, C. A. J., Honan, A. J., and Sappupo, J. S., 1983, “Aerodynamic
Platform Comparison for Jet Stream Electricity Generation,” J. Energy, 7, pp.
wt ⫽ tether weight per unit length 17–24.
WBA ⫽ weight of rotating arm: pivot B to point A 关4兴 Fletcher, C. A. J., 1983, “On the Rotary Wing Concept for Jet Stream Elec-
WCTR ⫽ counterweight tricity Generation,” J. Energy, 7, pp. 90–92.
关5兴 Riegler, G., Riedler, W., and Harvath, E., 1983, “Transformation of Wind
WDB ⫽ weight of rotating arm pivot B to point D Energy by a High Altitude Power Plant,” J. Energy, 7, pp. 92–94.
WF ⫽ weight of flywheel 关6兴 Goela, J. S., 1979, “Wind Power Through Kites,” Mech. Eng. 共Am. Soc.
WK ⫽ weight of kite Mech. Eng.兲, 42, pp. 42–43.
WLOAD ⫽ weight of load 关7兴 Goela, J. S., 1983, “Project Report II on Wind Energy Conversion Through
ⴱ Kites,” IIT Kanpur DST Project Report No. DST/ME共JSG兲/81-84/26/2.
WLOAD ⫽ normalized load weight= WLOAD / 0.5␳V2AK 关8兴 Varma, S. K., and Goela, J. S., 1982, “Effect of Wind Loading on the Design
zkite ⫽ kite elevation of a Kite Tether,” J. Energy, 6共5兲, pp. 342–343.
z0 ⫽ reference elevation 关9兴 Goela, J. S., Somu, N., Abedinzadeh, R., and Vijaykumar, R., 1985, “Wind
Loading Effects on a Catenary,” J. Wind. Eng. Ind. Aerodyn., 21, pp. 235–
␣ ⫽ geometric angle of attack of kite with respect 249.
to ground 关10兴 Goela, J. S., Vijaykumar, R., and Zimmermann, R. H., 1986, “Performance
␣asc ⫽ geometric angle of attack of kite during ascent Characteristics of a Kite-Powered Pump,” ASME J. Energy Resour. Technol.,
␣dsc ⫽ geometric angle of attack of kite during 108, pp. 188–193.
关11兴 Goela, J. S., 1984, “Final Report on Wind Energy Conversion Through Kites,”
descent IIT Kanpur DST Project Report No. DST/ME共JSG兲/81-84/26/2.
␣eff ⫽ effective angle of attack of kite, ␣eff = ␣ − ␾ 关12兴 Canale, M., Fagiano, L., Ippolito, M., and Milanese, M., 2006 “Control of
␣L=0 ⫽ kite angle of attack for zero lift Tethered Airfoils for a New Class of Wind Energy Generator,” Proceedings of
the 45th IEEE Conference Decision and Control, San Diego, CA, pp. 4020–
␤ ⫽ angle of kite motion with respect to horizontal 4026.
⌬x ⫽ retraction spring deflection 关13兴 Canale, M., Fagiano, L., and Milanese, M., 2007, “Power Kites for Wind
⌬t ⫽ numerical time step Energy Generation,” IEEE Control Syst. Mag., 27, pp. 25–38.
⌬tAOA ⫽ time period for angle of attack change 关14兴 Williams, P., Lansdorp, B., and Ockels, W., 2008, “Optimal Crosswind Towing
and Power Generation With Tethered Kites,” J. Guid. Control Dyn., 31共1兲, pp.
⌬zLOAD 2−4 ⫽ elevation change of load weight: engaged 81–93.
⌬zLOAD 4−2 ⫽ elevation change of load weight: disengaged 关15兴 Matos, C., Mahalingam, R., Ottinger, G., Klapper, J., Funk, R., and Komerath,
␥ ⫽ angle of rotating arm N., 1998, “Wind Tunnel Measurements of Parafoil Geometry and Aerodynam-
ics,” 36th AIAA Aerospace Sciences Meeting, Reno, NV, AIAA Paper No.
␥asc ⫽ rotating arm angle for ␣asc 98-0606.
␥dsc ⫽ rotating arm angle for ␣dsc 关16兴 White, F. M., 2008, Fluid Mechanics, 6th ed., McGraw-Hill, New York, Chap.
␪ ⫽ angle of kite tether with respect to horizontal 9, pp. 421–429.
关17兴 Kuethe, A. M., and Chou, C. Y., 1998, Foundations of Aerodynamics: Bases of
␾ ⫽ angle of local velocity vector with respect to Aerodynamic Design, 5th ed., Wiley, New York, Chap. 6, pp. 169–210.
horizontal 关18兴 Anderson, J. D., 2007, Fundamentals of Aerodynamics, 4th ed., McGraw-Hill,
␳ ⫽ air density New York, Chap. 5, pp. 391–463.

Journal of Solar Energy Engineering NOVEMBER 2010, Vol. 132 / 041009-11

Downloaded 03 Nov 2010 to 128.205.114.91. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

You might also like