You are on page 1of 6

Hypersonic Glider Control using Higher Order Sliding Mode Control

C. Tournes1, Senior Member, IEEE, G. Hanks2


AbstractHypersonic glider is a new class of missile
being developed. It characterized by sub-orbital
velocities and cruising altitudes in the 40-50 km altitude.
They pose a very challenging control problem in that
their aerodynamic model and eventual propulsion
models are largely uncertain. This paper proposes a
Higher Order Sliding mode control solution to the
problem, where the flight path angle rate and ground
track angles are controlled using the angle-of-attack and
roll angle as control parameter. Simulation results
present the flight profile and the wide reach footprint of
this type of vehicle. They show that the proposed control
solution achieves excellent results.
1.

INTRODUCTION

he idea of powered and un-powered hypersonic


vehicles cruising at sub-orbital velocities in the
upper stratosphere is not new, the idea has been
proposed as early as fifty years ago. There are reports that
TOPOL-M (SS-27) is actually a hypersonic glider. One of
the major control problems to be encountered with such a
vehicle is the scarcity of aerodynamic data, and the
likelihood of the model used for the guidance and automatic
pilot design to be largely uncertain. The problem is
compounded by lateral bank-to-turn maneuvers giving the
vehicle large lateral divert capabilities up to 1000 km on
each side of the trajectory.
A possible approach to achieve desired robustness is to
estimate in real-time part or entirety of the plant response
using Disturbance accommodating Observers. The system
response is governed by effects that are known and well
modeled, and by effects that are uncertain and unknown. The
Linear Adaptive control approach consists in estimating in
real-time the effects of unknown and uncertain terms and to
compensate in real-time for their effects. Disturbance effects
are the target motion and the effects of off-nominal values of
the missile response. To that effect we use Disturbance
Accommodating Control [1-3] theory to design
corresponding filters, and use the real-time estimation of the
effects of the disturbances to compensate their effects.
Another approach to overcome robustness issues is
Sliding Mode control (SMC). The main problem in using
SMC [4-6] is in the necessity of smoothing high frequency
switching control for the price of loosing robustness.
Higher order sliding mode control (HOSM) developed in
works [7-9] mitigates the problems associated with SMC,
i.e. HOSM is applicable to the systems with an arbitrary
relative degree, and continuous control functions can be
1

C. Tournes. Chief Scientist Davidson Technologies, Huntsville, AL 35806


USA (e-mail christianTournes@Davidson-Tech.com)
2
G. Hanks Program Manager Davidson Technologies, Huntsville, AL 35806
USA (e-mail greggHanks@Davidson-Tech.com)

978-1-4244-1884-8/08/$25.00 2008 IEEE

achieved while robustness is retained. The continuous 2sliding control, in particular super-twisting control [10-11]
provide high quasi continuous sliding mode control, are
robust to the disturbances but lack smoothness. Since in this
application we do not design the pitch autopilot it is possible
to use the super-twisting control. For future work including
the control of pitch motion, we will resort to a High Order
Smooth Sliding Mode Control providing the smoothness of
the angle-of-attack pseudo command such that it can be
tracked by inner-loop pitch autopilot. This will be done
using the works [12-14].
This paper simulates the flight of a hypersonic glider,
beginning by its acceleration by a launcher, its initial
unpowered ballistic segment, the controlled re-entry into the
atmosphere and its hypersonic glide, with or without lateral
turn maneuvers.
The paper is structured as follows: Section 2 presents the
formulation of the problem and governing equations. Section
3 presents the calculation of the launcher fly-out. Section 4
presents the design of the autopilot. Section 5 presents the
results obtained.
2.

FORMULATION OF THE PROBLEM

Hypersonic vehicle kinematics can be defined in Earth


Centered Inertial coordinates (ECI) coordinates [15-16]
simply as

w1
d
r = v = w2
dt
w3
w1
d F
Faero( m, h) r
3 ,
w2 = ( )I +
dt m
m
r
w3

(1)

where r, v represent ECI position and velocity and


I = T( , )v , v is the vector representing booster axis and
is T( , ) transformation matrix from velocity to body

axes, obtained by two successive rotations , around r, h


is the vector normal to the orbital plane and F represents the
magnitude of the thrust.
Note that during the first stage aerodynamic forces are
calculated, in velocity axes, the drag is aligned opposite to
w, the direction of the lift is obtained by calculating the
vector orthogonal to both the velocity vector and the body
axis and by a / 2 rotation of the velocity axis around this
vector. In this problem the nominal thrust magnitude is
constant during each stage; with the possibility of using
termination devices to achieve premature termination of the
current stage thrust. We will show that as result of the
robustness of sliding mode control, considerable

274

disturbances can be applied to the thrust magnitude without


any noticeable effect on trajectory accuracy.

atmosphere is modeled using Exponential Atmospheric


Model of Wertz [18].

We have at our disposal two parameters m, h defining


the desired orientation of the vehicle (initially the launch
vehicle and then the hypersonic glider) plus the possibility to
shut down the thrust of thr on-going stage. A system of three
axes i v , i h, i n oriented along the ECI velocity vector, the
specific angular momentum and the vector orthogonal to the
previous two such that the axis system is direct. The control
along the specific angular momentum is referred as the outof-plane or transverse control, and the control along the
third axis as the normal control or radial control.
Corresponding controls are u h , u n normalized to thrust
magnitude, and no control along the velocity vector can be
applied, other than the termination command. Commanded
orientation of the vehicle is defined as the unity vector

Remark 2: The Mach number is calculated using the


magnitude of the velocity with respect to Earth, the ECI
velocity- the rotation velocity of Earth at the latitude and
altitude.
Combining Eqs.(1), (4) and (5) one obtains during the
unpowered portion of the flight

w1
d
r = v = w2

dt
w3
w1 (cos()in sin()i y)0.7p(z)Sref M 2CL

d
+ Dist
(6)
=
w2
m

dt
w3

2
2
I*booster = u ni n + u hi h + 1 u n u h i v (2)

where iv , i h , i n are the unit vectors along velocity, specific


angular momentum and the third vector completing the
reference frame. The launcher dynamics are represented in
vector format, with each component of the launcher attitude
modeled as

i = booster ( i + u i )
Ibooster =

v Iv + h Ih + n In
2v + 2n + 2h

((3)

The advantage of the formulation if that I booster is directly


defined in ECI coordinates and the launcher kinematics can
be defined in ECI coordinates simply as per Eq. (1).
Remark 1: The normalized thrust along the velocity vector
is necessarily set to l = 1 u u , thus adequate
saturation bounds must be set to provide sufficient normal
and out of plane control authority while keeping the
component of thrust acceleration along the velocity large
enough, that is typically 1 > 0.6 .
2
n

2
h

3/2

Dist =

r iv0.7p(z)Sref M2(Cd0 + kCL )


r

(4)

The modeling of the hypersonic glider mission requires


solving the following problems.
1. Calculate ECI vector n orthogonal to the nominal orbital
plane.
2. Steer the launcher trajectory such as to achieve some
desired burn-out flight path angle.
3. Command the re-entry maneuver aimed at breaking down
the vertical re-entry velocity.
4. Calculate command angle-of-attack required to regulate a
constant flight path angle rate in the presence of eventual
roll maneuvers.
The equation governing the flight path angle can be obtained
by projecting Eq. (6) along i n as

 =

L( ) = 0.7 p ( z )S ref M CL
2

(5)

Where p(z) represents pressure function of altitude, M the


Mach number, and S ref the reference surface, and is the
angle of attack. Aerodynamic lift gradient CL is based on
slender body theory and drag coefficients use
k , Cd 0 results from slender body theory and Krasnovs
Aerodynamics of bodies of Revolution. [17]. The

mV

bal
V

(7)

v cos( )

bal (m s ) =
2
( R Earth + z ) R Earth + z
2

3/ 2

0.7 p( z ) S ref M 2CL

where

Drag D, and lift L are given respectively by


D( ) = 0.7 p( z ) S ref M 2(Cd 0 + kCL

Where , the angle of attack is the control and Dist is a


disturbance.

Aerodynamic effects were introduced as drag along the


velocity axis and lift orthogonal to it [15-16], as

F aero = L( ) cos( )i n D i v L( ) sin( )i y

3. CALCULATION OF THE LAUNCHER FLY-OUT


The vector normal to the orbital plane is obtained by
crossing out the ECI position vector at launch and ECI
position vector at arrival. The initial ECI launch is given by

275

xi G1cos( i) cos( Li)


yi = G1sin( i) cos( Li) ;
= 6378137 m

a Earth
zi G 2sin( Li)

a Earth
+ hi

2
2

G1 1 (2 f f )sin( Li)
2
; f = 1/ 298.3
G 2 =
a
Earth(1 f )

+ hi
1 (2 f f 2)sin( Li) 2

(8)

With i = Gi + SG , where Li and Gi are the launch


geodetic latitude and longitude and SG the sidereal angle at
launch time. Parameters G1 and G2 are used to model Earth
oblatness and hi is the height at launch.
Similarly the final ECI position is given by

G1cos( f ) cos( Lf )
= G1sin( f ) cos( Lf )

G 2 sin( Lf )

a Earth
+ hf

2
2

G1 1 (2 f f )sin( Lf )
2

G 2 =
a
Earth (1 f )

+ hf
1 (2 f f 2 )sin( Lf ) 2

xf
yf

zf

h=

4. DESIGN OF THE AUTOPILOT


A first open loop maneuver is conducted to break the
vertical velocity as the hypersonic glider re-enters the
atmosphere. The altitude at which this maneuver is initiated
is given by

z break

bo > 0.2 130, 000 m

= 0.2 > bo 0 40, 000 + 450, 000

bo = 0 40, 000 m

bo

m (11)

The amplitude of command acceleration is given by

x i xf
yi yf

z i zf

bo > 0.2 20

( m s ) = 0.2 > bo 0 5 + bal + 750 bo (12)

bo = 0 bal + 5

The flight path angle is stabilized at a constant


value * = 0.0075 rad. The sliding proportional integral
surface n that, being stabilized at zero, provides for the
desired compensated tracking error dynamics is chosen as
n = ( * )d + ( * ) ; n = 30 rad / (normalized time) (13)

(10)

The booster fly-out is calculated as follows:


A terminal flight path angle is chosen. All the stages
except the first one are flown with a linear angle of attack
law, with a maximum value Max at the beginning of the
second stage and a null value at burn out.
1. A first guess of terminal altitude and velocity is made
2. An iterative process is initiated, where:
2.1 The flight is integrated backwards from the burn out
condition until the beginning of the second stage.
2.2 The first stage is flown forward with a flight path angle
constant = / 2 , then time = t turn the launcher flight path
angle is rotated linearly to a value

4. Iteration process stops when the differences between the


end of the first stage and the beginning of second stage are
quasi identical.
5. The burn out altitude can be modified, by modifying the
value of Max . The burn-out velocity can be modified by
modifying the amount of fuel.

(9)

With f = Gf + SG + EarthT , where Lf and Gf are


arrival geodetic latitude and longitude and SG the sidereal
angle at launch time, hf is the height at arrival, T the flight
time and Earth is Earths rotation rate. The orbital plane is
orthogonal to

x i xf
yi yf

z i zf

2.3 Comparison of range, altitude, time, velocity magnitude


and flight path angle between the forward integration of first
stage flight with backwards integration of upper stages,
exhibits differences
3. Burn-out range, time, velocity magnitude and first stage
beg are modified and a new iteration is conducted.

beg in 4 seconds, then a

normal gravity turn is applied until the end of the first stage.

If we differentiate Eq. (13) with respect to time we see that


according to Eq. (7) the control appears, and the SISO
dynamics of the surface can be written as

 n = g (t ) + bu ,

(14)

which will be further interpreted as the sliding variable


dynamics calculated along the system trajectory. The
condition n = 0 defines the system motion on the sliding
surface, u is a control input that needs to be
continuous, and g (t ) is a sufficiently smooth uncertain
function. The problem that is addressed in this section is to
design continuous control u that drives ,  0 in finite
time.

276

Equation (13) shows that, by achieving n = 0 , the tracking


error e = * converges to zero asymptotically according
to the eigenvalue of Eq. (13). Equation (13) can be robustly
stabilized at zero using traditional sliding mode control [5],
which exhibits a high frequency switching. The problem is
that the launcher attitude needs to be flipped-on with a
frequency, usually as large as 10Hz which is impossible to
achieve in practice.
Since in this application we are only required to generate a
continuous control law, we suggest using a second order
sliding mode control in particular a super-twisting algorithm
that generates a continuous control function and, which is in
the same time completely robust to the effects of the
disturbances.
Remark 3: When the control needs to be designed not only
to be continuous but also smooth, which would have been
the case if we had included a nested angle-of-attack
autopilot, we use a smooth second order sliding mode
controller with a built in sliding mode observer as in [19-20].
The uncertain function written as

g (t ) = * + *

0.7 p ( z ) S ref M 2CL 


mV

<L

(15)

normally small, and with the prescribed flight path angle


constant glide path law * = 0 normally small.
The super-twist control is designed in terms of

= 100 n

1/2

sign ( n ) + 20 sign ( n )d

(16)

Commanded angle of attack is finally given by

com =

100 n

1/2

sign ( n ) + 20 sign ( n )d + bal

0.7 p( z )S ref M 2CL cos( )

(17)

5. RESULTS

A two stage launcher solid booster burning in two


minutes was used to achieve burn out flight path angles in
the 0-0.02 rad range and burn-out velocities in the 64006850 m/s range. Not surprisingly, the range achievable
varies considerably. The example used to illustrate the
concept was obtained for a burn-out flight path angle =
0.01rad and a burn-out velocity = 6400 m/s. The flight
profile is shown in Fig. 1

Is evidently bounded, since the uncertainty on the lift


coefficient gradient is bounded, the angle of attack rate is

Fig. 1 Flight Profile

277

Fig. 2 Sliding Surface


Interestingly, several roll maneuvers are conducted during
the constant glide segment with roll angles (in velocity
axes) of 30, 45 and 60 degrees. The result of the roll
maneuver is to increase the value of com as per Eq. (14)
the effect of which is to augment the induced drag and
thus cause a greater longitudinal deceleration. It is
remarkable to notice on Fig. 1 that the four altitudes vs.
range profiles are perfectly superimposed which
demonstrated the robustness of the autopilot. The
autopilot drives the sliding surface Eq. (12) very rapidly
to zero, and the sliding surface remains near zero
thereafter as illustrated, Fig. 2.

Another interesting result is the lateral divert maneuver.


The roll maneuver achieves lateral maneuvers with
amplitudes of up to 1000 km in each side. This means that
such hypersonic glider can each any point over an area of
4000 km long and 2000 km wide. The defense against
such weapon has no means for predicting the point aimed
until the maneuver is initiated.
Claims have been made that the TOPOL-M can start
scram-jets during its atmospheric glide, allowing the
missile to re-exit the atmosphere, which of course would
make the prediction of its target even more problematic.

Fig. 4 Velocity magnitude


Fig. 3 Lateral divert
278

Lateral maneuvers impose a penalty in that they reduce


considerably the longitudinal velocity as shown in Fig. 4. It
shows that when no lateral maneuver is conducted
longitudinal velocity of 6200 m/s is attained at a range of
4760 km (in Fig. 3) and at time from launch of 841 sec.
When a 60 deg roll maneuver is conducted, the longitudinal
velocity can drop to 2300 m/s in less than four minutes.
CONCLUSION

We have presented an evaluation of an intercontinental


hypersonic glider. It shows that even without powered
maneuvers this type of weapon has a very wide reaching
domain. A simple angle-of-attack autopilot based on second
order sliding mode control could perfectly regulate the
prescribed flight path angle. The control design is simple
and requires very limited information on the vehicle model.
Owing to the property of invariance of sliding mode control
to matched disturbances, considerable disturbance can be
introduced without any effect on the controller.
REFERENCES

[1] C. Tournes, and C.D. Johnson, Direct-Lift Design


Strategy for longitudinal Control of Hypersonic Aircraft
using Subspace Stabilization Control Techniques,
Proceedings of the 1999 AIAA Guidance Navigation
and Control Conference, Reston VA.
[2] Johnson, C.D., A New Approach to Adaptive Control,
chapter in the book Advances in Control and Dynamic, ,
Vol. 27, Leondes, C.T., (ed.) Academic Press, NY,
1988, chapter 1.
[3] Johnson, C.D., Example Application of the Linear
Adaptive Control Technique, International Journal of
Adaptive Control and Signal Processing, Vol. 3, John
Wiley & Sons, Ltd., NJ, 1989, pp. 111-119
[4] Edwards, C., and Spurgeon, S., Sliding Mode Control,
Taylor & Francis, Bristol, PA, 1998.
[5] Utkin, V., Guldner, J., and Shi, J., Sliding Modes in
Electromechanical Systems, Taylor and Francis,
London, 1999.
[6] Zhou, D., Mu, C., and Xu, W., Adaptive Sliding-Mode
Guidance of a Homing Missile. Journal of Guidance,
Control and Dynamics, 22, 4, 1999, pp. 589-594.
[7] Davila, J., Fridman, L.M., and Levant, A., Second-order
sliding-mode observer for mechanical systems, IEEE
Transactions of Automatic Control, 50 (11), 2005, pp.
1785-1789.
[8] Floquet, T., Barbot, J.P., and Perruquetti., W., Higherorder sliding mode stabilization for a class of nonholonomic perturbed systems. Automatica 39, pp.
2003, 1077 1083.
[9] Orlov, Y., Finite time stability and robust control
synthesis f uncertain switched systems. SIAM Journal

on Control and Optimization, 43 (4), 2005, pp. 12531271.


[10] Levant, A., Robust exact differentiation via sliding
mode technique. Automatica, 34, 3, pp. 379-384, 1998.
11] Levant, A., Quasi-continuous high-order sliding-mode
controllers. IEEE Transactions on Automatic Control,
50, 11, 2005, pp. 1812 1816.
[12] Bhatt, S., and Bernstein. D., Finite Time Stability of
Continuous Autonomous Systems, SIAM Journal on
Control and Optimization, Vol. 38, No. 3, 2000, pp.
751-766.
13] Bacciotti, A., and Rosier, L., Liapunov Functions and
Stability in Control Theory, Lecture notes in control
and information sciences, 267, Springer-Verlag, NewYork, 2001.
[14] Levant., A., Homogeneity approach to high-order
sliding mode design, Automatica, 41, (5) 2005.pp. 823830,
[15] Norton, M., Spacecraft Navigation and Guidance,
Springer, London GB, 1998, pp. 41-54.
[16] Tournes, C., Shtessel, Y.B., Predictive Launcher
Guidance using Second Order Sliding Mode Control
Proceedings of the 2005 AIAA Guidance Navigation
and Control, Reston, VA.
[17] Krasnov, N.F., Aerodynamics of Bodies of
Revolution, American Elsevier Publishing Company,
Inc, New York, 1970.
[18] Vallado, D., Fundamentals of Astrodynamics and
Applications, Second Edition, Microcosm Press, El
Secundo, CA, 2001, p. 537.
[19] Levant, A., Homogeneity approach to high-order
sliding mode design, Automatica , 41, (5) pp. 823-830.
[20] Shtessel, Y., Shkolnikov, I., Levant, A., Smooth
Second Order Sliding Modes: Missile Guidance
Application, Automatica, 2007

279

You might also like