Sunteți pe pagina 1din 26

AIAA Atmospheric Flight Mechanics Conference and Exhibit AIAA 2006-6130

21 - 24 August 2006, Keystone, Colorado

Modeling and Simulation of 9-DOF Parafoil-Payload


System Flight Dynamics

Om Prakash∗and N. Ananthkrishnan†
Indian Institute of Technology - Bombay, Mumbai 400076, India

Parafoil-payload system requires a 9-DOF dynamic model representing three degrees


of freedom each for rotational motion of the canopy, and payload, and three degrees of
freedom for translational motion of the confluence point of the lines. The gliding flight of
parafoil-payload system is directly affected by choice of rigging angle. The glide angle is
controlled by symmetric trailing edge (brake) deflection and turn is effected by asymmetric
brake deflection. The parafoil trim and stability characteristics are a complex function of
rigging angle and magnitude of downward deflection of left and right brakes. The present
work uses a 9-DOF model to analyze gliding and turning flight of the parafoil-payload
system for different choices of design rigging angle subjected to change in left and right
brake deflections. Also, 9-DOF model is used to show the possibility of heading control
using Nonlinear Dynamic Inversion law by close-loop simulations

Nomenclature
A, B, C apparent mass terms
b canopy span
d length of brake control line pulled
CD drag coefficient
CL lift coefficient
CY side force coefficient
Clp , Clr rolling moment damping coefficients
Cmc/4 pitching moment coefficient at quarter chord
Cmq pitching moment damping coefficient
Cnr , Cnp yawing moment damping coefficients
c canopy chord length
F force
g acceleration due to gravity
IA , IB , IC apparent inertia terms
IF parafoil apparent inertia matrix
M mass matrix
m mass
q̄ dynamic pressure (= 12 ρV 2 )
p, q, r roll, pitch, and yaw rates
Sb payload (body) cross-section area
Sp parafoil planform area
Tb transformation matrix from inertial reference frame to payload reference frame
Tp transformation matrix from inertial reference frame to parafoil reference frame
t canopy maximum thickness
u, v, w velocity components along reference frame
V total velocity
X, Y, Z body-fixed reference frame
∗ Ph.D. Student, Department of Aerospace Engineering; omp@aero.iitb.ac.in. Student Member AIAA.
† Associate Professor, Department of Aerospace Engineering; akn@aero.iitb.ac.in. Senior Member AIAA.

1 of 26

American Institute of Aeronautics and Astronautics

Copyright © 2006 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
x, y, z position components
α angle of attack
β sideslip angle
γ glide angle
δa dimensionless asymmetric deflection (= d/c)
δs symmetric brake deflection angle
µ rigging angle
ρ air density
φ, ψ, θ Euler roll, yaw, and pitch angles, respectively
Subscript
b referred to payload (body)
c referred to link joint C
e referred to earth fixed
p referred to parafoil (canopy)
pa from parafoil CG to AC in parafoil frame
cb from joint C to payload CG in joint C frame
cp from joint C to parafoil CG in joint C frame
pI referred to air inside the canopy
β stability derivative with respect to β
δa control derivative with respect to δa
Superscript
b referred to payload body
l referred to lines
p referred to parafoil canopy
Abbreviations
AC Aerodynamic Center
CG Center of Gravity
DOF Degree of Freedom
N DI Nonlinear Dynamic Inversion

I. Introduction
Parafoils find wide use in UAV (Unmanned Aerial Vehicle), CRV (Crew Return Vehicle), GPADS (Guided
Parafoil Air Drop System) to sports activities due to their good gliding as well as control characteristics.
The glide angle is controlled by symmetric trailing edge (brake) deflection and turn is effected by asymmetric
brake deflection. The canopy open leading edge, large number of lines, and payload of arbitrary shape are the
drag producing components. Also, deflection of trailing edges (brakes) for parafoil control results in increase
in drag. The presence of vertical offsets in centers of various aerodynamic forces from overall system center
of gravity gives rise to nonlinear trim and stability characteristics of the parafoil-payload system. Parafoil
geometric parameters like rigging angle, have a strong effect on the trim and stability characteristics of the
system.
Although the parafoil canopy has very small rigid mass as compared to payload mass, the included
air mass and apparent mass result in total parafoil canopy mass being comparable to payload mass. The
additional included air mass and apparent mass have a large effect on the rotational motion of parafoil
canopy. Due to the presence of a confluence point of the line connecting the payload to the parafoil, parafoil
and payload exhibit independent rotational motion. Hence, the parafoil-payload system dynamics is required
to be modeled as a two-body problem. Thus, the parafoil-payload system requires a 9-DOF dynamic model
representing three degrees of freedom each for rotational motion of the canopy and payload, and three degrees
of freedom for translational motion of the confluence point.
Slegers and Costello1 used 9-DOF model to investigate control issues for a parafoil-payload system with
left and right parafoil brakes used as the control mechanism. They were able to show that parafoil-payload
system can exhibit two basic modes of directional control, namely, roll steering and skid steering. Mooij et
al.2 presented a 9-DOF flight dynamic model of parafoil-payload system which they used to develop a flight
simulation environment for the Small Parafoil Autonomous Delivery System (SPADES). Machin el al.3 used

2 of 26

American Institute of Aeronautics and Astronautics


a two-body 8-DOF flight dynamic model to determine the aerodynamic characteristics of parafoil recovery
system used for safely landing a crew return vehicle. Heise and Muller4 used a nonlinear high-fidelity two-
body 8-DOF model (considering 6-DOF motion of parafoil and 2-DOF relative motion of payload) to develop
a unified software tool for the modeling, simulation, and highly realistic visualization of parafoil system.
Iosilevskii5 used standard static stability analysis to show that most forward CG position of gliding
parachute results in loss of longitudinal stability. Lingard6 has carried out longitudinal and lateral stability
analysis of parafoil-payload system showing effect of rigging angle and brake deflection. Brown7 illustrated
the effect of scale and wing loading on turn response of a parafoil-payload system using closed-form turn
equation. Crimi8 presented a lateral stability analysis of gliding parachutes. He determined the effect of
parafoil anhedral angle, suspension line length, and glide slope in spiral divergence and oscillatory response.
Machin et al.3 investigated a range of rigging angles from 4 to 16 deg through actual parafoil flight tests,
with most of the testing focused on 10 and 13 deg, to get required trim glide performance. They reported
that a parafoil rigged at higher rigging angle has higher turn performance.
Bifurcation methods are a convenient tool for numerical analysis of trim and stability of dynamical
systems, including nonlinear effects. Bifurcation methods were introduced to flight dynamics by Carroll and
Mehra,9 and Zagainov and Goman.10 Bifurcation methods have been used to study complete high alpha
dynamics, and constrained flight maneuvers, such as level trims.11 A previous study by us12 used bifurcation
methods to show the possibility of multiple trims for a given rigging angle and symmetric brake deflection
of a parafoil-payload system using a longitudinal 4-DOF model.
The present work uses a 9-DOF model to analyze effect of choice of rigging angle, and use of left and right
brake deflections on the gliding and turning flight of parafoil-payload system. The gliding flight analysis
uses bifurcation methods to determine effect of rigging angle and symmetric brake deflection on trim and
stability of the system. Using best glide rigging angle obtained from gliding analysis, turning flight of the
parafoil-payload system is analyzed for different left and right brake deflections.

II. Parafoil-Payload 9-DOF model

Cmc/4 Xp
p a αp
γ

V
Yp
d
Zp
µ

C Xc
Yc
b
Xb
Zc

Yb Zb

Figure 1. 9-DOF parafoil-payload system

As shown in Fig. 1, the parafoil-payload system is modeled as a fixed-shape parafoil canopy of mass
mp , and a payload body of mass mb . The mass centers of canopy and payload are connected to joint C
through rigid massless links. Both the parafoil and the payload are free to rotate about joint C, but are
constrained by the internal joint force (F xc , F yc , F zc ) at C. The 9-DOF motion of parafoil-payload system
is described by three inertial position components of joint C (xc , yc , zc ), as well as three Euler orientation

3 of 26

American Institute of Aeronautics and Astronautics


angles of the parafoil (φp , θp , ψp ) and payload (φb , θb , ψb ). Formulation of equations of motion requires three
reference frames, namely, parafoil reference frame (Xp , Yp , Zp ) fixed to parafoil CG, payload reference frame
(Xb , Yb , Zb ) fixed to payload CG, and joint C fixed reference frame (Xc , Yc , Zc ) parallel to inertial Earth-fixed
frame (Xe , Ye , Ze ). For the present study, the parafoil mass center is assumed to be at the parafoil canopy
mid-baseline point. Then, rigging angle µ is defined as the angle between the parafoil link and the parafoil
Zp axis. The rigging angle µ = 0 indicates that the forwardmost and rearmost parafoil lines, meeting at
joint C, are of equal length. The 9-DOF parafoil-payload model is similar to that in Slegers and Costello,1
except that we do not employ the spring and damper modeling of relative yawing motion between parafoil
and payload due to the lines. In our model, the payload is not reoriented when the parafoil yaws during a
turn as the payload aerodynamic and gravitational forces are independent of payload orientation.

A. Equations of Motion
The kinematic equations for parafoil and payload are given as :
     

 ẋe  
  ẋc    uc 

ẏe = ẏc = vc (1)

  
   
  

że żc wc

    

 φ̇b  1 Sφb tθb Cφb tθb   pb 

 
θ̇b =  0 Cφb −Sφb  qb (2)

 
 
 

ψ̇b 0 Sφb /Cθb Cφb /Cθb rb

    

 φ̇p 
 1 Sφp tθp Cφp tθp   pp 

 
θ̇p =  0 Cφp −Sφp  qp (3)

 
 
 

ψ̇p 0 Sφp /Cθp Cφp /Cθp rp

The common shorthand notation for trigonometric function is employed, where sin α ≡ Sα , cos α ≡ Cα ,
and tan α ≡ tα . The 9-DOF model of combined parafoil canopy and payload system in matrix form is
represented as :
    
−Mb Rb 0 Mb Tb Tb Ω̇b B1
    
 0 −(Mp + MF )Rcp (Mp + MF )Tp −Tp   Ω̇p   B2 
   =   (4)
 Ib 0 0 −Rcb Tb   V̇c   B3 
0 Ip + IF 0 Rcp Tp Fc B4

B1 = FbA + FbG − Ωb × Mb Ωb × Rcb


B2 = FpA + FpG − Ωp × (Mp + MF )Ωp × Rcp + MF Ωp × Tp Vc − Ωp × MF Tp Vc
B3 = −Ωb × Ib Ωb
B4 = MA
p − Ωp × (Ip + IM )Ωp

where,
   
0 −rb qb 0 −rp qp
   
Ωb × =  rb 0 −pb  ; Ωp × =  rp 0 −pp  (5)
−qb pb 0 −qp pp 0
and
         

 pb 
 
 pp 
 
 xcb 
 
 xcp 
 
 uc 

Ωb = qb ; Ωp = qp ; Rcb = ycb ; Rcp = ycp ; Vc = vc (6)

 
 
 
 
 
 
 
 
 

rb rp zcb zcp wc

4 of 26

American Institute of Aeronautics and Astronautics


The velocity vector and joint force vector at joint C are Vc = [uc , vc , wc ]T and Fc = [F xc , F yc , F zc ]T
respectively. The aerodynamic force and weight force vector, respectively, of payload are :
   

 ub  
 −Sθb 

A 1 b G
Fb = ρVb Sb CD vb ; Fb = mb g Sφb Cθb (7)
2 
 
 
 

wb Cφb Cθb
b
where only drag force CD is modeled for payload.
The aerodynamic force and moment at parafoil CG, and parafoil weight force vector are :
   

 CX  
 bCl 

A
FpA = q̄p Sp CY ; M p = q̄ S
p p cCmc/4 + xpa CZ (8)

 
 
 

CZ bCn
 

 −Sθp 

FpG = mp g Sφp Cθp (9)

 

Cφp Cθp
The parafoil apparent mass and moment of inertia matrices are :
   
A 0 0 IA 0 0
   
MF =  0 B 0 ; IF =  0 IB 0  (10)
0 0 C 0 0 IC
The matrix Tb represents the transformation matrix from an inertial reference frame to the payload body
reference frame :
 
Cθb Cψb Cθb Sψb −Sθb
 
Tb =  Sφb Sθb Cψb − Cφb Sψb Sφb Sθb Sψb + Cφb Cψb Sφb Cθb  (11)
Cφb Sθb Cψb + Sφb Sψb Cφb Sθb Sψb − Sφb Cψb Cφb Cθb
The matrix Tp represents the transformation matrix from an inertial reference frame to the parafoil body
reference frame :
 
Cθp Cψp Cθp Sψp −Sθp
 
Tp =  Sφp Sθp Cψp − Cφp Sψp Sφp Sθp Sψp + Cφp Cψp Sφp Cθp  (12)
Cφp Sθp Cψp + Sφp Sψp Cφp Sθp Sψp − Sφp Cψp Cφp Cθp
The payload and parafoil mass center velocity components in payload and parafoil frame respectively are :
     

 ub   
 ẋc  
 xcb 

vb = Tb ẏc + Ωb × ycb (13)

 
 
 
 
 

wb żc zcb
     

 up 
 
 ẋc 
 
 xcp 

vp = Tp ẏc + Ωb × ycp (14)

 
 
 
 
 

wp żc zcp
The apparent mass and inertia terms are based on the following formulas given by Lissaman and Brown7 :
A = ρ 0.913πt2b/4
B = ρ 0.339πt2c/4
C = ρ 0.771πc2b/4
IA = ρ 0.630πc2b3 /48
IB = ρ 0.872 4c4 b/48π
IC = ρ 1.044πt2b3 /48 (15)

5 of 26

American Institute of Aeronautics and Astronautics


The mass mpI = ρbct/2 and moment of inertia of the included air in the parafoil canopy also need to be
added to the apparent mass and inertia terms, respectively.

B. Rigging Angle Modeling


The rigging angle µ is the angle between the line joining mid-baseline point of the canopy to joint C and the
line parallel to Zp axis passing through the mid-baseline point. Therefore,

zcp = Rp cos µ
xcp = Rp sin µ
ycp = 0 (16)

Also, zcb = Rb , xcb = 0 and ycb = 0.

C. Aerodynamic Model
Parafoil aerodynamic force and moment coefficients are modeled as:

CL = CL (αp , δs ) + CLδa δa
p p
CD = CD (αp , δs ) + CDδa δa
CY = CYβ β + CYr (rp b/2Vp ) + CYδa δa

p
CX = (−CD up + CL wp )/Vp
p
CZ = (−CD wp − CL up )/Vp

pp b rp b
Cl = Clβ β + Clp + Clr + Clδa δa
2Vp 2Vp
qp c
Cm = Cmc/4 (αp , δs ) + Cmq + Cmδa δa
2Vp
pp b rp b
Cn = Cnβ β + Cnp + Cnr + Cnδa δa (17)
2Vp 2Vp

Here symmetric brake deflection δs corresponds to zero, half and full brake conditions, while asymmetric
brake deflection δa is defined as δa = d/c with value 0.0-0.24, where d is length of brake control lines pulled
down.
The magnitude of total velocity vector in payload and parafoil reference frame is :
1
Vb = (u2b + vb2 + wb2 ) 2
1
Vp = (u2p + vp2 + wp2 ) 2 (18)

The payload and parafoil angles of attack, parafoil sideslip angle, and glide angle are computed, respec-
tively, as

αb = tan−1 (wb /ub )


αp = tan−1 (wp /up )
βp = sin−1 (vp /Vp )
γ = sin−1 (cos αp cos βp sin θp − sin βp sin φp cos θp − sin αp cos βp cos φp cos θp ) (19)

III. Geometric and Aerodynamic Data


A. Parafoil-payload Geometry
The parafoil and payload geometric parameters, as given in Table 1, are taken from Lingard6 except where
indicated otherwise.

6 of 26

American Institute of Aeronautics and Astronautics


Table 1. Parafoil-payload system geometry.

Parameter Value Parameter Value


c 3.75 m mb 135 kg
t 0.18 c Rb 0.5 m
b 7.5 m Sb 0.5 m2
mp 5 kg Sp 28 m2
mPI ρbct/2 xpa 0.25c
Rp -7.5 m Cmqp -1.864 ((cb/2Vp )−1 )7

B. Longitudinal Aerodynamic Coefficients


p
The longitudinal aerodynamic characteristics consists of parafoil aerodynamic drag force coefficient CD , lift
force coefficient CL , and pitching moment coefficient at quarter chord Cmc/4 . These are highly nonlinear
function of angle of attack αp for zero, half and full symmetric brake deflections. These aerodynamic
p
coefficients as shown in Figs. 2, are taken from Lingard.6 CD and Cmc/4 are corrected for line drag CD l
l
acting at middle of line length using relation CD = nRp d/Sp , where n = 40 is number of lines and diameter
of a line, d = 35 mm.

C. Lateral Aerodynamic Derivatives


The parafoil lateral aerodynamic characteristics consisting of lateral stability and control derivatives are
taken from Brown7 for a similar parafoil-payload system. These are listed in Table 2.

Table 2. Lateral Derivatives.

Parameter Value Parameter Value


CYβ -0.0095 /deg CLδa 0.2350
CYr -0.0060 (rb/2Vp )−1 CDδa 0.0957
Clβ -0.0014 /deg CYδa 0.1368
Clp -0.1330 (rb/2Vp )−1 Cmδa 0.2940
Clr 0.0100 (rb/2Vp )−1 Clδa -0.0063
Cnβ 0.0005 /deg Cnδa 0.0155
Cnp -0.0130 (rb/2Vp )−1
Cnr -0.0350 (rb/2Vp )−1

IV. Gliding Flights


The gliding flight of parafoil-payload system is directly affected by choice of rigging angle and magnitude
of symmetric brake deflection applied. Bifurcation method is used to analyze gliding trim and stability
characteristics of parafoil-payload system subject to symmetric zero, half, and full brake, for different design
rigging angles. For a best glide rigging angle, effect of dynamic change of symmetric brake on gliding
characteristics of parafoil-payload system is investigated using 9-DOF simulation.
The present bifurcation analysis makes use of 9-DOF dynamic model of parafoil-payload system unlike the
longitudinal 4-DOF model used earlier.12 The 9-DOF kinematic and dynamic equations of parafoil-payload
system from Eqs. (2), (3), and (4) are represented as:

χ̇ = f (χ, U ) (20)

7 of 26

American Institute of Aeronautics and Astronautics


where χ = [φb , θb , ψb , φp , θp , ψp , pb , qb , rb , pp , qp , rp , uc , vc , wc ] and U = [δs , µ], the δs is symmetric brake
deflection for zero, half and full brake configuration, and µ is rigging angle.
Using the AUTO200013 continuation algorithm, trims for gliding flight of parafoil-payload system are
computed with zero, half, and full symmetric brakes, and different rigging angles. The trims so computed
are shown in terms of bifurcation diagrams of parafoil angle of attack αp , and glide angle γ, in Figs. 3 and
4, for zero, half, and full brake. The trims represented by solid lines are stable, and those represented by
dashed lines are unstable. Points of onset of instabilities are bifurcation points represented by solid square.
The bifurcation diagrams show possibility of multiple trims and regions of unstable gliding flight, as observed
in earlier work.12
The bifurcation diagram in Fig. 3 shows two regions of glide instabilities for zero, half, and full brake
conditions. In zero brake condition, a very small region of instability occurs at α = 12 − 15 deg and a large
region of instability occurs between α = 25 deg to α = 32 deg. For the full symmetric brake, a small region
of instability between α = 8 deg to α = 15 deg and a large instability region between α = 20 deg to α = 30
deg. The discontinuous bifurcation diagram (due to numerical problems) for symmetric half brake deflection
shows near about same trim and stability characteristics as for zero brake deflection indicating no effect of
brake on system glide upto half brake deflection.
In Fig. 4 bifurcation diagram for γ shows parafoil-payload system has minimum stable glide angle at
rigging angle µ = 9 deg for zero and half symmetric brake. At this rigging angle, full brake shows large glide
angle γ = 55 deg. At rigging angles sufficiently larger than µ = 9 deg, full brake results in two stable trim
glide angles, one showing γ ≈ 30 deg and the other at γ ≈ 50 deg. At rigging angles sufficiently smaller
than γ = 9, zero brake results in two stable trim glide angles, one showing γ ≈ 28 deg and the other at
γ ≈ 50 deg. Thus, rigging angle µ = 9 deg is the optimum value of rigging angle to obtain good glide as well
as good flare characteristics (large γ under full brake conditions) for parafoil-payload system. The rigging
angle µ = 9 deg is selected for further studies.

A. Simulation of Full Symmetric Brake


Bifurcation analysis only gives trims and stability of the dynamical system. To predict the transient response,
one needs to carry out time simulation. This is especially important in case of large perturbations and for
large and/or rapid control inputs. Numerical simulations of dynamics of parafoil-payload system have been
carried out using the 9-DOF dynamic model.
We investigate the effect of applying full brake on the dynamics of parafoil-payload system with rigging
angle µ = 9 deg. A typical maneuver is simulated which is shown in Figs. 5 and 6 with the following sequence
of control inputs. The system is trimmed at the zero brake condition, then a ramp input over a time of 1 sec
is applied from zero brake to full brake starting at t = 5 sec. The full brake input is held fixed until t = 30
sec, when a ramp input over a time of 1 sec is applied to recover from full brake to zero brake condition.
When parafoil brake is deflected from zero to full starting at t = 5 sec, the system takes 1 − 2 cycles of
oscillations over a period of about 8 sec before it reaches a post-stall trim state. This trim state has a steep
glide angle γ = 52 deg, and high αp = 38 deg. This is the post-stall trim state computed in bifurcation
diagram of Figs. 3 and 4 for µ = 9 deg. Thus, for full brake deflection, the parafoil system glides with
canopy-stalled configuration. Effectively, with full brake deflection, the parafoil-payload system descends
much like a conventional circular parachute with L/D ratio less than 1.
When the parafoil canopy reverts back to zero brake from full brake deflection starting t = 30 sec, the
system takes at least 3 cycles of oscillation over more than 20 sec to successfully recover to the original zero
brake trim.

B. Simulation of Different Choice of Rigging Angles


We investigate the effect of three different choices of rigging angles on dynamic maneuver as well as stall
recovery. The simulation is started with initial condition as parafoil-payload system in trim glide with zero
brake deflection and drop altitude 1000 m.
The simulation results are presented in Fig. 7 for three choices of rigging angles, i.e., µ = 6, 9 and 12
deg, and discussed below. As shown in bifurcation diagrams, Figs. 3 to 4, µ = 6 deg corresponds to parafoil-
payload system in post-stall zero brake and post-stall full brake trim, µ = 9 deg corresponds to pre-stall
(minimum glide slope) zero brake and post-stall full brake trim, and µ = 12 deg corresponds to pre-stall zero
brake and pre-stall full brake trim.

8 of 26

American Institute of Aeronautics and Astronautics


Flare maneuver : On applying symmetric brake at t = 5 sec, the dynamics is characterized by parafoil
pitch oscillation coupled with pendulum effect of payload.
The rigging angle µ = 6 deg achieves highest αp but least damped oscillation. For µ = 6 and 9 deg, the
system shows large reduction in forward (horizontal) velocity and large negative glide slope, because system
finds post-stall trims at these rigging angles. On the other hand, for µ = 12 deg, the system shows smaller
decrease in forward velocity and minor increase in negative glide slope as the system finds only pre-stall trim
in full brake condition. After 2.5 sec of the ramp input initiation at t = 5 sec, the system reaches maximum
pitch up angle. The full brake is usually used for dynamic flare maneuver, therefore, it is the peak transient
that is important, not, the trim value. On application of the ramp input, at the point of maximum positive
pitch up, rigging angle µ = 6 deg shows maximum reduction in forward velocity to uc ≈ 4 m/s, whereas
µ = 9 and 12 deg show reduced forward velocity of uc ≈ 6.5 m/s and uc ≈ 8 m/s, respectively. Hence, the
smaller rigging angle µ = 6 deg is most promising to obtain good flare maneuver, followed by µ = 9, deg
which is only marginally worse and also corresponds to minimum glide slope for zero brake flight.
Stall recovery: As the rigging angles µ = 6 and 9 deg for full brake condition put the parafoil-payload
system in post-stall trim, a ramp input over 1 sec at t = 30 sec is applied to symmetric brake from full to zero
to check for stall recovery. This results in stall recovery for µ = 9 deg with lightly damped phugoid oscillation,
but, the system with µ = 6 deg is not able to recover from stall due to the post-stall trim characteristics at
this rigging angle. Whereas parafoil-payload system with µ = 12 deg does not show post-stall trim in full
brake, so there is no need for stall recovery.
Thus, although the rigging angle µ = 6 deg looks most promising for flare maneuver, but due to inability
of stall recovery on redeploying zero brake it is not advisable for practical purpose. Therefore, the rigging
angle µ = 9 deg which shows adequate flare maneuver and good stall recovery, as well best glide slope, is
advisable for practical purpose.

V. Turning Flight
Parafoils are capable of performing circular turns using either asymmetric brake alone (i.e., either left or
right brake deflection), or asymmetric in combination with symmetric brake deflection (i.e., differential brake
deflection). Unlike aircraft which obtain turn rate due to the lift component created by banking, parafoils
achieve required turn rate due to differential drag force generated by asymmetric brake deflection on left
and right canopy. The asymmetric brake not only causes differential drag but affects lift force too, thereby
altering L/D ratio and glide slope of the system. Thus, asymmetric deflection of left and right parafoil
braking not only affects lateral-directional (turn and bank) response but also longitudinal response of the
parafoil-payload system.
The parafoil circular turn maneuver is accompanied by loss of height while turning for a sufficiently large
time interval. Therefore, parafoil turn maneuver is performed in order to lose excess height over the landing
zone so as to reach the appropriate height to execute the flare maneuver. A larger turn rate gives a small
circular turn radius which helps the system to stay close to the landing zone.

A. Full Asymmetric Brake


The 9-DOF parafoil-payload model derived in chapter 3 is used to investigate the turn performance using
different deflections of right and left brake. The system is rigged at µ = 9 deg, which gave optimum glide
performance in the previous chapter. Full right and left brake deflections correspond to δa = +0.24 and
δa = −0.24, respectively.
The simulation for turning flight of parafoil-payload is carried out separately for full right and full left
brake deflections and results are shown in Figs. 8, 9 and 10. The system is allowed to glide with zero left
and zero right brake deflection for 25 sec. Then, full right or left brake, as appropriate, is deflected over
an interval of 1 sec and remains deflected thereafter. The system loses about 100 m height and covers a
distance of about 275 m range before asymmetric brake deflection is initiated, it then takes approximately
two cycles of lateral oscillations over 10 sec to achieve a trimmed turning flight.

1. Figure 8 shows trajectories of the system for left and right full brakes. The system descends executing
perfect symmetric left and right circular turns for the full left and right brake deflections, respectively.
Right brake deflection causes parafoil to turn towards right showing positive cross range and perform
circular turns. This is because right brake down gives increased drag on right side of the canopy

9 of 26

American Institute of Aeronautics and Astronautics


generating positive yawing moment leading to right turn. The left brake shows similar turning response
towards left side showing negative cross range. The system loses around 450 m height in approximately
three and a half circular turns of 75 m radius for both left as well as right brake deflection.
2. For the two cases of deflection, Fig. 9 shows symmetric parafoil yaw and bank response. The sideslip
angle is typically small. The symmetric left and right turn response is expected due to symmetric
lateral aero data used in the 9-DOF dynamic model. The large angle of attack trim on application of
full asymmetric brake shows that the canopy is stalled during turn leading to low L/D, hence steeper
glide slope. Thus, the system descends faster during turning with full asymmetric brake. This could be
helpful when parafoil-payload systems are required to execute circular turns in order to reduce excess
height over landing zone.
3. Figure 10 shows reduced horizontal velocity, increased vertical velocity, and cross velocity of opposite
signs at joint C for right and left full brakes. Note that the joint C reference frame and payload reference
frame do not reorient during turning flight, therefore, components of joint forces Fcx , Fcy exchange
horizontal joint force among themselves showing false cyclic oscillations. For the same reason, payload
outward swing motion during turning flight is exchanged between pitch and bank angles showing false
oscillation. The small payload yawing angle build up is due to false pitch and bank angles and must
be ignored. As the aerodynamics of payload is independent of bank and yaw angles, it does not affect
the dynamics of the overall system.

B. Various Right Brakes


Figure 11 shows turning flight simulation for various right brake deflections (positive asymmetric brake
deflections), i.e., δa = 0.05, 0.12 and 0.24 keeping δs = 0.

1. As compared to δa = +0.24 (full) asymmetric brake, parafoil-payload system with δa = +0.12 (half)
asymmetric brake shows approximately half the yaw rate, bank angle and sideslip angle. This is
expected as the lateral aerodynamics is modeled as a linear function of δa (i.e., no angle of attack
dependency). Thus, system executes a turn double the radius in turning flight with half brake. The
longitudinal trims show post-stall angle of attack for both full and half asymmetric brake. Thus, the
glide slope with half brake is only a little less than for full brake, and the system loses just a little less
height as with full brake during 360 deg turn.
2. As compared to δa = +0.24 (full), +0.12 (half) asymmetric brakes, the small asymmetric deflection
case, i.e., δa = 0.05, shows parafoil angle of attack in pre-stall region during turn. Thus, system turns
with glide angle and descent rate nearly same as zero brake deflection.

Thus, parafoil-payload system with smaller asymmetric brake deflection remains in pre-stall angle of
attack region giving a shallow glide slope, and hence descends slower taking larger turn radius. On the
other hand, system with large symmetric deflections trim post-stall angles of attack giving steeper glide
slopes, and hence descend faster with smaller turn radius. Therefore, parafoil-payload system can make use
of large asymmetric brake deflection to descend faster while staying nearby in turn of small radius over the
landing zone. This is useful during excess height loss phase before touchdown. But, at small symmetric
brake deflection, turn rate is very small, and it takes a large turn radius, consequently a large time interval
is needed to complete one turn. This is useful for sport hovering over a region for longer time.

VI. Closed-Loop Simulations


Autonomous parafoil-payload system is required to deliver payload to a desired landing point from an
arbitrary release point using GPS-based technologies. For the large part of the autonomous flight phase,
parafoil-payload system is expected to mainly maintain desired heading angle leading to landing zone (LZ).
After reaching the LZ, the parafoil-payload system is required to lose the excess height by flying in S-turns
or circular turns.
The basic requirement of a control law for a flight vehicle like parafoil-payload system is that it should
be able to attain and hold the vehicle in a desired attitude and heading using available controls, in this
case, left and right brakes. Also, if the parafoil-payload system encounters undesirable flight conditions, the

10 of 26

American Institute of Aeronautics and Astronautics


controller should be able to bring it back to some desirable safe situation. Bifurcation and simulation results
show that parafoil-payload system may be expected to exhibit high-angle-of-attack, nonlinear dynamics, for
a typical choices of given combination of control input and rigging angle.
Nonlinear Dynamic Inversion (NDI) is a method to access closed-loop characteristics of a system prior
to design and testing of a control law. NDI method does not require to linearize the vehicle dynamics, and
can be used to develop closed-loop system dynamics even at nonlinear aerodynamic regimes. Thus, NDI is
used to design simple generic nonlinear control law for the whole region without detailed analysis. In recent
times, NDI-based control law has been suggested as effective means for flight control problems, e.g. Steer14
has applied NDI flight control method to a second generation supersonic aircraft, and Mulder et al.15 have
designed the flight controller for a re-entry vehicle using NDI.

A. Nonlinear Dynamic Inversion (NDI) Theory


NDI is based on the concept of canceling the original dynamics of the system and replacing it with the
desired dynamics designed by the designer. This is achieved by inverting the governing equations of the
system dynamics based on the measured system states and input commands. The forces and the moments
are modeled as functions of measured state variables. Opposing forces and moments are then commanded
to negate the undesired contributions (forces and moments acting on the body) followed by the application
of the desired forces and moments needed to accomplish the maneuver. The net result is that the output
should perfectly match the input, if the inversion is exact and actuator saturation and rate limits are ignored.
Usually dynamic inversion requires equal number of inputs and states. If the number of inputs are less than
the available states, then all the states cannot be controlled. The states which remain uncontrolled, then
behave according to the nature of the open-loop dynamics. Multiple time-scale using inner loop inversion
for fast system dynamics and outer loop inversion for slow system dynamics allows to control more number
of states for a given control input. The dynamic inversion has been used for aircraft longitudinal as well as
lateral control at high angle of attack.

B. NDI Applied to Parafoil Turning Flight


The 9-DOF, nonlinear dynamic model with complete nonlinear longitudinal aerodynamics from -10 deg to
+80 deg angle of attack is used to implement the NDI for heading control of parafoil-payload system. The
parafoil-payload system is assumed to be in trimmed gliding flight with symmetric brake deflection, δs , kept
zero and asymmetric brake command, δa , used as the variable control for the forward simulation. The
asymmetric brake deflection, δa , is used to command parafoil turn angle, ψp . The external command is the
required heading angle ψpc , which is required to be converted into an asymmetric brake deflection demand δad
in the inversion procedure. This is achieved through a two-step process consisting of an outer loop inversion
and an inner loop inversion as shown in Fig. 12. The outer loop calculates the desired parafoil yaw rate rpc
based on the parafoil turn command ψpc . The inner loop then calculates the control demand, δad based on
the yaw rate rpc calculated in the outer loop. Note that pp , qp loops are not closed, hence the qp , rp obtained
due to rpc are accepted.
In order to develop the NDI law, the 9-DOF equations of motion of parafoil-payload system from Eq. (4)
are represented as:

AẊ = {B0 + Bδa δa } (21)


where
 
FbA + FbG − Ωb × Mb Ωb × Rcb
 A 
 Fp0 + FpG − Ωp × (Mp + MF )Ωp × Rcp + MF Ωp × Tp Vc − Ωp × MF Tp Vc 
B0 =  
 −Ωb × Ib Ωb 
MA p0 − Ωp × (Ip + IM )Ωp

 p

{−CD (αp , δs )up + CLp (αp , δs )wp }/Vp
A  
Fp0 = q̄p Sp  CY β β + CY r rp 2Vb p ;
p p
{−CD (αp , δs )wp − CL (αp , δs )up }/Vp

11 of 26

American Institute of Aeronautics and Astronautics


 
b{Clβ β + Clp pp 2Vb p + Clr rp 2Vb p }
 
MA
p0 = q̄p Sp  c 
 c{Cmc/4 (αp , δs ) + xpa CZ + Cmq qp 2Vp } 
b{Cnβ β + Cnp pp 2Vb p + Cnr rp 2Vb p }

and
 
03×1
 A 
 Fpδ 
Bδa =  a

 03×1 
MA pδa

   
(−CDδa up + CLδa wp )/Vp bClδa
   
A
Fpδ a
= q̄p Sp  CY δa ; MA
pδa = q̄p Sp  cCmδa 
(−CDδa wp − CLδa up )/Vp bCnδa

C. Inner Loop Inversion


The equation of motion for dynamics of yaw rate is

ṙp = a(12) {B0 + Bδa δa } (22)

where a(12) is a row vector corresponding to 12th row of matrix A−1 in Eq. (21).
The yaw rate command, rpc , and the feedback, rp are used to calculate the error er = (rpc − rp ), and the
desired yaw acceleration ṙpd is modeled as,

ṙpd = ωr (rpc − rp ) (23)

where ωr is the bandwidth.


The inversion function, by manipulation of Eq. (22), leads to the following expression:

ṙpd − a(12) B0
δac = (24)
a(12) Bδa

This is the asymmetric brake deflection required to achieve the demanded yaw acceleration, ṙpd . Substituting
Eq. (24) in Eq. (22) for δa , we get

ṙp = ωr (rpc − rp ) (25)

Thus, the yaw rate dynamics has been converted into a first-order system with time constant T = 1/ωr . If
(rpc − rp ) is denoted as ∆rp and if rpc is a constant command (i.e., ṙpc = 0), we can get

∆ṙp = −ωrp ∆rp (26)

D. Outer Loop Inversion


The outer loop produces yaw rate command rpc by inversion of the ψ˙p equation. The design of the inversion
law for the slow state, ψp , is based on the assumption that the fast state, rp responds much faster than the
slow state ψp . It is assumed that the transient dynamics of the fast state occurs so quickly that it has a
negligible effect on ψp .
The Equation for dynamics of yaw angle is given by third row of Eq. (3) as:

sin φp cos φp
ψ˙p = qp + rp (27)
cos θp cos θp

12 of 26

American Institute of Aeronautics and Astronautics


When the commanded parafoil yaw angle, ψpc , is compared with the feedback, ψp , we get ψ̇pd as

ψ̇pd = ωψ (ψpc − ψp ) (28)

where ωψp is the bandwidth. The inversion function derived after manipulation of Eq. (27) is as follows:
   
sin φp cos φp
rpc = ψ̇pd − qp / (29)
cos θp cos θp

The calculated yaw rate is fed as the command to the inner loop. In order to obtain the yaw rate
instantly, the inner loop is attributed a higher bandwidth, i.e. ωr = 5 whereas ωψ = 2, hence faster response
as compared to outer loop.
ω
If the inversion is exact and ωψr ∼ 0, it is equivalent to substituting rpc in Eq. (27), giving the equation:

ψ̇p = ψ̇pd = ωψp (ψpc − ψp ) (30)

If ψpc is a constant command (i.e., ψ̇p = 0), then Eq. (30) becomes

∆ψ̇p = −ωψp ∆ψp (31)

where ∆ψ˙p = (ψpc − ψp ).


Equation (31) suggests that the outer-loop inversion is that of a first-order system with time constant
T = 1/ωψ . However in practice, the inner and outer loops couple, ans the obtained dynamics is not precisely
as given in Eqs. (25) and (30).

E. Turn Response using Inner Loop Alone


Effectiveness of NDI inner loop controller to generate required turn rate rp is analyzed using 9-DOF simula-
tion. First, the simulation is carried out without NDI for full asymmetric deflection δa = 0.24, which shows
steady yaw rate of 15.6 deg/sec. Then corresponding yaw rate is given as input yaw rate command, i.e.,
rpc = 15.6 deg/sec in cosed loop using NDI inner loop only. In both cases the input command in the form of
ramp input for 1 second is initiated at 25 sec of simulation and kept fixed for all time. The results obtained
from both the simulations are compared in Fig. 13.
The NDI inner loop is able to generate commanded yaw rate efficiently with only a minor error in steady
state in all the variables. As the NDI inner loop works by inverting the r˙p equation alone, the other states
of the system respond according to their open loop dynamics. Thus, although the NDI inversion ought to
result in linear response of the inverted dynamics, due to coupling with other states, the yaw rate dynamics
also responds to the dynamics of the other states. Nevertheless, using NDI, the system not only closely
achieves the commanded turn rate, but also, the corresponding longitudinal post-stall characteristics seen in
the open-loop simulations. The radius of turn for closed-loop is little bigger than open-loop with little lag
as seen from height vs range plot.

F. Heading Tracking using both Loops


Using NDI with inner as well as outer inversion loops, the parafoil-payload system should be able to track
the commanded heading. There is no control on the pitch and roll attitudes or rate variables, therefore
they will be free to respond according to their open-loop dynamics. Figure 14 shows how perfectly the
NDI-controller is able to track a series of heading command, i.e., ψp by generating required rp and hence
asymmetric brake deflection δa . The other state variables respond according to their own natural dynamics.
The system is required to undergo four turning commands followed by straight gliding flight. During each
turn the system undergoes post-stall angle of attack flight and subsequently recovers from stall when input
command is withdrawn. The large variations in lateral-directional states is due to momentarily large brake
deflection input generated by NDI law. Whereas, longitudinal states show large variation due to post-stall
angle of attack flight corresponding to large brake deflections.

13 of 26

American Institute of Aeronautics and Astronautics


VII. Conclusions
We have been successful in employing a bifurcation and continuation algorithm to compute all trim states
and their stability with varying rigging angle for three different symmetric brake settings. The results are
all plotted on a single diagram making it easy to draw conclusions about the parafoil dynamic behavior. For
instance, it is clearly seen that the best glide angle is obtained for an angle of attack within 1-2 deg of the
stall angle of attack, when rigged correctly. Also, the glide ratio is not appreciably affected for symmetric
brake deflection up to 50 per cent. For larger than optimum rigging angle, the post-stall trim is at a relatively
lower angle of attack and glide ratio, thereby giving poor flare maneuver. All these conclusions match with
observations made in the literature from previous experiences with designing parafoil systems.
Through simulations, we have demonstrated that the choice of rigging angle from our bifurcation studies
is indeed the correct one. All maneuvers, trims, glides, flare, and turns, are seen to be cleanly carried out for
the chosen best rigging angle. For a lower value of rigging angle, our simulations clearly show poorer glide
and inability to recover from post-stall trims on removal of full brake deflection.
An NDI control strategy has been proposed and successfully implemented in closed-loop simulations.
Our simulations with the NDI control law are able to cleanly track yaw rate and heading angle commands.
In fact, it is clear from our simulations that a properly designed and rigged parafoil system needs only the
yaw loops to be closed, the pitch and roll response show good behavior in open loop including post-stall
flight phases and recovery.

References
1 Slegers, N., and Costello, M., “Aspects of Control for a Parafoil and Payload System,” Journal of Guidance, Control and

Dynamics, Vol. 26, No. 6, 2003, pp. 898-905.


2 Mooij, E., Wijnands, Q.G.J. and Schat, B., “9-dof Parafoil/Payload Simulator Development and Validation,” AIAA

Modeling and Simulation Technologies Conference and Exhibit, Austin, TX, August 11-14, 2003.
3 Machin, R. A., Iacomini, C. S, Cerimele, C. J., and Stein, J. M. “Flight Testing the Parachute System for the Space

Station Crew Return Vehicle,” Journal of Aircraft, Vol. 38, No. 5, 2001, pp. 786-799.
4 Heise, M., and Muller, S., “Dynamic Modeling and Visualization of Multi-Body Flexible Systems,” AIAA Modeling and

Simulation Technologies Conference and Exhibit, Providence, Rhode Island, August, 2004.
5 Iosilevskii, G., “Center of Gravity and Minimal Lift Coefficient Limits of a Gliding Parachute,” Journal of Aircraft, Vol.

32, No. 6, 1995, pp. 1297-1302.


6 Lingard, J. S., “The Performance and Design of Ram-Air Parachutes,” Precision Aerial Delivery Seminar, Technical

Report, Royal Aircraft Establishment, Aug 1981.


7 Brown, G. J., “Parafoil Steady Turn Response to Control Input,” 12th AIAA Aerodynamic Decelerator Systems Tech-

nology Conference and Seminar, London, UK, May 10-13 1993, pp. 248-254.
8 Crimi, P., “Lateral Stability of Gliding Parachute,” Journal of Guidance, Control and Dynamics, Vol. 5, No. 5, 1982,

pp. 529-536.
9 Carroll, J. V., and Mehra, R. K., “Bifurcation Analysis of Nonlinear Aircraft Dynamics,” Journal of Guidance, Control

and Dynamics, Vol. 5, No. 5, 1982, pp. 529-536.


10 Zagainov, G. I., and Goman, M. G., “Bifurcation Analysis of Critical Aircraft Flight Regimes,” International Council of

the Aeronautical Sciences, September 1984, pp. 217-223.


11 Ananthkrishnan, N., and Sinha, N. K., “Level Flight Trim and Stability Analysis using Extended Bifurcation and Con-

tinuation Procedure,” Journal of Guidance, Control and Dynamics, Vol. 24, No. 6, 2001, pp. 1225-1228
12 Prakash, O., Daftary A. and Ananthkrishnan, N, “Trim and Stability Analysis of Parafoil/Payload System using Bifur-

cation Methods,” 18th AIAA Aerodynamic Decelerator Systems Technology Conference and Seminar, Munich, Germany, May
2005.
13 Doedel, E. J., Paffenroth, R. C. Champneys, A.R., Fairgrieve, T.F., Kuznetsov, Y.A., Sandstede, B., and Xang, X.,

“AUTO2000: Continuation and Bifurcation Software for Ordinary Differential Equations (with Hom Cont),” Technical Re-
port,California Inst. of Technology, Pasadena, CA, USA, 2001.
14 Steer, A. J., “Application of NDI Flight Control to a Second Generation Supersonic Transport Aircraft,” Computing and

Control Engineering Journal, June 2001, pp. 108-114.


15 daCosta, R. R., Chu, Q. P., and Mulder, J. A., “Reentry Flight Controller Design Using Nonlinear Dynamic Inversion,”

Journal of Spacecraft and Rockets, Vol. 40, No. 1, Jan-Feb 2003, pp. 64-71.

14 of 26

American Institute of Aeronautics and Astronautics


1.4 1
. Zero Brake
1.2
* Half Brake 0.8
1 o Full Brake
p
0.6
CD 0.8
CL
0.6 0.4
0.4
0.2
0.2

0 0
−20 0 20 40 60 80 −20 0 20 40 60 80
Angle of Attack, α Angle of Attack, αp
p

0 3.5

−0.1 3

−0.2 2.5

−0.3 2
L/D
Cm
c/4−0.4 1.5

−0.5 1

−0.6 0.5

−0.7 0
−20 0 20 40 60 80 −20 0 20 40 60 80

Angle of Attack, α Angle of Attack, αp


p

Figure 2. Longitudinal aerodynamics data

15 of 26

American Institute of Aeronautics and Astronautics


80.

70.
Half Brake
60.

50. Full Brake

40.
αp

30.

20. Zero Brake

10.
Half Brake
0.
-10. -5. 0. 5. 10. 15.
Rigging Angle, µ

Figure 3. Bifurcation diagram of the parafoil angle of attack (Full line: stable trim, dashed line: unstable
trim, solid square: Hopf bifurcation point)

80.

70.
Half Brake
Full Brake
60.

50.
γ
40.
Zero Brake
30.

20.
Half Brake

10.
-10. -5. 0. 5. 10. 15.
Rigging Angle, µ

Figure 4. Bifurcation diagram of glide angle of parafoil-payload system (Full line: stable trim, dashed line:
unstable trim, solid square: Hopf bifurcation point)

16 of 26

American Institute of Aeronautics and Astronautics


1 20

0.5
0
φp, deg

θ , deg
0

p
−20
−0.5

−1 −40
0 20 40 60 0 20 40 60
Time, sec Time, sec
1 40

0.5 30

α , deg
rp

0 20

p
−0.5 10

−1 0
0 20 40 60 0 20 40 60
Time, sec Time, sec
1 0

0.5
−20
βp, deg

γ, deg

0
−40
−0.5

−1 −60
0 20 40 60 0 20 40 60
Time, sec Time, sec
1 1000
Cross Range m

0.5
Height, m

900
0
800
−0.5

−1 700
0 200 400 600 0 200 400 600
Range, m Range, m
1 100

0.5
δa, deg

δs, deg

0 50

−0.5

−1 0
0 20 40 60 0 20 40 60
Time, sec Time, sec

Figure 5. Dynamic effect of full symmetric brake δs deflections with µ = 9 deg parafoil canopy variables

17 of 26

American Institute of Aeronautics and Astronautics


15 500

10 0
uc, m/s

,N
cx
F
5 −500

0 −1000
0 20 40 60 0 20 40 60
Time in sec Time in sec
1 1

0.5 0.5
vc, m/s

Fcy , N
0 0

−0.5 −0.5

−1 −1
0 20 40 60 0 20 40 60
Time in sec Time in sec
8 1800

6 1600
wc, m/s

,N

4 1400
cz
F

2 1200

0 1000
0 20 40 60 0 20 40 60
Time in sec Time in sec
1 20

0.5
0
φb, deg

θb, deg

0
−20
−0.5

−1 −40
0 20 40 60 0 20 40 60
Time in sec Time in sec
1 1

0.5 0.5
ψb, deg

ψp, deg

0 0

−0.5 −0.5

−1 −1
0 20 40 60 0 20 40 60
Time in sec Time in sec

Figure 6. Dynamic effect of full symmetric brake δs deflections with µ = 9 deg connection point and payload
variables including parafoil yaw angle

18 of 26

American Institute of Aeronautics and Astronautics


30 10

20 0
10
θ , deg

−10

θ , deg
0
b

p
−20
−10

−20 −30

−30 −40
0 10 20 30 40 50 0 10 20 30 40 50
Time in sec Time in sec
1 60

50
0.5
40

α , deg
δa

Time in sec
0 30

p
20
−0.5
10

−1 0
0 10 20 30 40 50 0 10 20 30 40 50
Time in sec Time in sec
100 0

80
−20
60
δ , deg

−40
γ, deg
s

40
−60
20

0 −80
0 10 20 30 40 50 0 10 20 30 40 50
Time in sec Time in sec
14 10

12 8
Time in sec
10
u , m/s

w , m/s

6
8
c

4
6

4 2

2 0
0 10 20 30 40 50 0 10 20 30 40 50
Time in sec
Time in sec

Figure 7. Dynamic effect of full brake for different rigging angles Solid line: 12 deg, Thick solid line: 9 deg, and
Dashed line: 6 deg. (All angles in deg)

19 of 26

American Institute of Aeronautics and Astronautics


1000

900
60
Height, H

800
40
700

20
600

0
0 Cross Range, Yc
50 −20
100
150 −40
Forward Range, Xc 200
250 −60
300

Figure 8. Turning flight for full left, and full right brake deflections with µ = 9 deg. Solid line: δa = +0.24;
Dotted line: δa = −0.24

20 of 26

American Institute of Aeronautics and Astronautics


20 0

10
−5
φ , deg

θp, deg
0
p

−10
−10

−20 −15
0 20 40 60 80 100 0 20 40 60 80 100
Time, sec Time, sec
20 20

10
15

α , deg
r , deg

p
p

10
−10

−20 5
0 20 40 60 80 100 0 20 40 60 80 100
Time, sec Time, sec
4 −15

2 −20
β , deg

γ, deg

0 −25
p

−2 −30

−4 −35
0 20 40 60 80 100 0 20 40 60 80 100
Time, sec Time, sec
1000 100
Cross Range m

50
Height, m

800
0
600
−50

400 −100
0 100 200 300 400 0 100 200 300 400
Range, m Range, m
1 0.4

0.5 0.2
a
s

0 0
δ

−0.5 −0.2

−1 −0.4
0 20 40 60 80 100 0 20 40 60 80 100

Time, sec Time, sec

Figure 9. Turning flight for full left, and full right brake deflections with µ = 9 deg. Solid line: δa = +0.24;
Dashed line: δa = −0.24, parafoil canopy variables

21 of 26

American Institute of Aeronautics and Astronautics


11 500

10
u , m/s

Fx
0
c

8 −500
0 20 40 60 80 100 0 20 40 60 80 100
Time in sec Time in sec
5 500
v , m/s

Fy
0 0
c

−5 −500
0 20 40 60 80 100 0 20 40 60 80 100
Time in sec Time in sec
6 1350

1300
5
w , m/s

Fz

1250
c

4
1200

3 1150
0 20 40 60 80 100 0 20 40 60 80 100
Time in sec Time in sec
20 20

10 10
φb, deg

θ , deg

0 0
b

−10 −10

−20 −20
0 20 40 60 80 100 0 20 40 60 80 100
Time in sec Time in sec
100 2000

50 1000
ψ , deg

ψp, deg

0 0
b

−50 −1000

−100 −2000
0 20 40 60 80 100 0 20 40 60 80 100

Time in sec Time in sec

Figure 10. Turning flight for full left, and full right brake deflections with µ = 9 deg. Solid line: δa = +0.24;
Dashed line: δa = −0.24, connection point and payload variables including parafoil yaw angle

22 of 26

American Institute of Aeronautics and Astronautics


20 0

15
−5
φ , deg

θp, deg
10
p

−10
5

0 −15
0 50 100 150 0 50 100 150
Time, sec Time, sec
20 20

15
15

αp, deg
p

10
r

10
5

0 5
0 50 100 150 0 50 100 150
Time, sec Time, sec
4 −15

−20
2
β , deg

γ, deg

−25
p

0
−30

−2 −35
0 50 100 150 0 50 100 150
Time, sec Time, sec
400 1000
Cross Range m

300 800
Height, m

200 600

100 400

0 200
0 100 200 300 400 500 0 100 200 300 400 500
Range, m Range, m
0.4 1

0.3 0.5
δa, deg

δs, deg

0.2 0

0.1 −0.5

0 −1
0 50 100 150 0 50 100 150
Time, sec Time, sec

Figure 11. Turning flight for various asymmetric brake deflections with µ = 9 deg. (Solid line: δa = +0.05;
Dashed line: δa = +0.12, Thick solid line: δa = 0.24.)

23 of 26

American Institute of Aeronautics and Astronautics


American Institute of Aeronautics and Astronautics

Figure 12. Block diagram of NDI law

δs= 0

Outer−Loop NDI Inner−Loop 9−DOF


. . Model
24 of 26

ψ cp Outer−Loop ψ dp Outer−Loop r cp Inner−Loop r dp Inner−Loop δa


d
X
Bandwidth Inversion Bandwidth Inversion
+ ωψ Dynamics + ωr Dynamics
− rp
− ψp
20 −2
input: δ =0.24
a
15 −4 input: rp =15.6 deg/s

−6

θ , deg
10
−8
φ , deg

p
5
−10
p

0 −12

−5 −14
0 20 40 60 80 100 0 20 40 60 80 100
Time in sec Time in sec
20 20

15 18

10 16
r , deg

5 14

α , deg
p

0 12

p
−5 10

−10 8
0 20 40 60 80 100 0 20 40 60 80 100
Time in sec Time in sec
4 −15

3
−20
β , deg

2
γ, deg

−25
p

1
−30
0

−1 −35
0 20 40 60 80 100 0 20 40 60 80 100
Time in sec Time in sec
1000 80

900
Cross Range m

60
800
Height m

40
700
20
600

500 0
0 100 200 300 400 0 100 200 300 400
Range m

Range m

Figure 13. Turn rate response of parafoil-payload system (Solid line: δac = 0.24; Dashed line: rpc = 15.66 deg/s).

25 of 26

American Institute of Aeronautics and Astronautics


20 0

10

θ , deg
−5
φ , deg

p
p

−10
−10

−20 −15
0 50 100 150 0 50 100 150
Time in sec Time in sec
20 20

10
15

α , deg
r . deg

p
p

10
−10

−20 5
0 50 100 150 0 50 100 150
Time in sec Time in sec
4 −10

2
−20
β , deg

γ, deg

0
p

−30
−2

−4 −40
0 50 100 150 0 50 100 150
Time in sec Time in sec
0.4 100

0.2 50
ψ , deg
δ , deg

0 0
p
a

−0.2 −50

−0.4 −100
0 50 100 150 0 50 100 150
Time in sec Time in sec
1000 400
Cross Range, m

300
Height, m

800
200
600
100

400 0
0 200 400 600 800 0 200 400 600 800

Range, m Range, m

Figure 14. Tracking capability of closed-loop parafoil-payload system (Dashed line: commanded yaw angle ψpc ;
Solid line: obtained yaw angle ψp ).

26 of 26

American Institute of Aeronautics and Astronautics

S-ar putea să vă placă și