You are on page 1of 13

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/220337692

Exponential Stabilization of the Inertia Wheel Pendulum Using Dynamic


Surface Control.

Article  in  Journal of Circuits, Systems and Computers · February 2007


DOI: 10.1142/S0218126607003514 · Source: DBLP

CITATIONS READS

16 456

4 authors, including:

N. Qaiser Naeem Iqbal Awan


Pakistan Institute of Engineering and Applied Sciences Pakistan Institute of Engineering and Applied Sciences
22 PUBLICATIONS   111 CITATIONS    67 PUBLICATIONS   421 CITATIONS   

SEE PROFILE SEE PROFILE

Amir Hussain
Edinburgh Napier University
559 PUBLICATIONS   10,565 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Cognitive Big Data Informatics (CogBID) Research Lab View project

Field Classification View project

All content following this page was uploaded by Amir Hussain on 27 December 2014.

The user has requested enhancement of the downloaded file.


May 15, 2007 7:33 WSPC/123-JCSC 00351

Journal of Circuits, Systems, and Computers


Vol. 16, No. 1 (2007) 81–92
c World Scientific Publishing Company

EXPONENTIAL STABILIZATION OF THE INERTIA WHEEL


PENDULUM USING DYNAMIC SURFACE CONTROL

NADEEM QAISER∗ and NAEEM IQBAL†


Department of Electrical Engineering,
Pakistan Institute of Engineering and Applied Sciences, Nilore,
Islamabad 44000, Pakistan
∗nadeem.qaiser@pieas.edu.pk
†naeem.iqbal@pieas.edu.pk

AMIR HUSSAIN
Department of Computer Science and Mathematics,
University of Sterling, Scotland, UK
a.hussain@cs.stir.ac.uk

NAEEM QAISER
Department of Computer and Information Sciences,
Pakistan Institute of Engineering and Applied Sciences, Nilore,
Islamabad 44000, Pakistan
naeemqaiser@pieas.edu.pk

Revised 7 November 2006

This paper considers the stabilization problem of Inertia Wheel Pendulum, a widely
studied benchmark nonlinear system. It is a classical example of a flat underactuated
mechanical system, for which the design of control law becomes a challenging task owing
to its underactuated nature. A novel nonlinear controller design, fusing the recently
introduced Dynamic Surface Control and the Control Lyapunov Function method, is
presented as the solution. Stability is analyzed using concepts from Singular Perturba-
tion Theory. The proposed design procedure is shown to be simpler and more intuitive
than existing designs. Advantages over conventional Energy Shaping and Backstepping
controllers are analyzed theoretically and verified using numerical simulations.

Keywords: Inertia wheel pendulum; underactuated mechanical systems; backstepping;


dynamic surface control; singular perturbation model.

1. Introduction
Inertia Wheel Pendulum (IWP), first introduced by Spong et al.,1 is a benchmark
nonlinear underactuated mechanical system. This is a mechanical control system
with fewer actuators (i.e., controls) than configuration variables. IWP has been

81
May 15, 2007 7:33 WSPC/123-JCSC 00351

82 N. Qaiser et al.

a test bed mainly for Energy Shaping and Damping Injection based approaches.
Spong et al.1 use a supervisory hybrid/switching control strategy to asymptotic
stabilization of the inertia wheel pendulum around its upright equilibrium point.
First, a passivity-based controller2 swings up the pendulum. Then, a balancing
controller obtained by Jacobian linearization or (local) exact feedback linearization
stabilizes the pendulum around its upright position.
Global stabilization of the IWP system can also be achieved using the Integrator
Backstepping procedure (IBS). Integrator backstepping, based on results obtained
by Sontag and Sussman,3 is a powerful step-by-step design tool. However it not
only suffers the problem of “explosion of terms” but also requires certain system
functions to be C n , see Ref. 4. The control law obtained through a cumbersome
design procedure is usually very complicated. Multiple Sliding Surfaces4 (MSS)
control, a procedure similar to integrator backstepping, avoids this phenomenon
but falls short of integrator backstepping in terms of theoretical rigor, as the need
for analytical differentiation is pushed to a numerical one.
Concept of Dynamic Surface Control (DSC), a dynamic extension to MSS, intro-
duced by Swaroop et al.5 resolves these issues by using low pass filters. It not only
addresses the issue of “explosion of terms” associated with integrator backstepping
but also solves the problem of finding the ith state reference (desired) trajectory
derivatives numerically for the MSS scheme. It is simple, more intuitive, and applies
to a more general class of systems as compared to the MSS scheme in Won and
Hedrick,5 as the requirement on the nonlinear function is to be C 1 only.
Dynamic Surface Control is exploited to stabilize the IWP, demonstrating the
design method simplicity that also results in a less complicated control law. The
design fuses the DSC technique and the Control Lyapunov Function method (CLF).
The design procedure is simpler than IBS and yields a control law that can be
implemented easily. Controller architecture is also simple, as it does not require
a supervisory controller like the one by Spong et al.1 However, stability analysis
require special attention as DSC is not used in regular fashion.
The paper starts formally with Sec. 2 containing the dynamic model for the
inertia wheel pendulum. Here necessary coordinate transformations are also given,
as the dynamic model is not in a control design amenable form. Controller design
strategy and procedure appear in Sec. 3, which also hosts detailed stability dis-
cussion. Section 4 presents simulation results comparing controller performance to
existing designs followed by brief concluding remarks in Sec. 5.

2. Dynamical Model
The IWP as illustrated in Fig. 1 is a planar inverted pendulum with a rotating
wheel at the end. The pendulum is to be controlled by the coupling torque generated
through wheel rotation, as the joint on the base is unactuated. The controller task is
to stabilize the pendulum in its upright equilibrium position while the wheel stops
rotating. The specific angle of rotation of the wheel is not important.
May 15, 2007 7:33 WSPC/123-JCSC 00351

Exponential Stabilization of the IWP Using Dynamic Surface Control 83

I2 q2
τ
m2
I1 ,L1

q1 m1 g

Fig. 1. The inertia wheel pendulum system.

The dynamical model of IWP can be easily obtained by using the Euler–
Lagrange method.6 Using configuration variables as shown in Fig. 1 the Lagrangian
is given as
1 T
L(q, q̇) = q̇ M q̇ − V (q1 ) , (1)
2
where (q, q̇) are generalized coordinates and quasi-velocities, M , the inertia matrix,
and V (q1 ) is the potential energy function given as1

V (q1 ) = w cos(q1 ) , where w := (m1 l1 + m2 L1 )g ,


d ∂L ∂L
− = Q(q)u
dt ∂ q̇ ∂q
yields the equations of motion given as
      
m11 m12 q̈1 − w sin(q1 ) 0
+ = , (2)
m21 m22 q̈2 0 τ

where m11 = m1 l12 + m2 L21 + I1 + I2 , m12 = m21 = m22 = I2 , and

I1 , I2 — moment of inertia of the pendulum and the wheel (kgm2 )


m1 , m2 — mass of the pendulum and the wheel (kg)
L 1 , l1 — length of the pendulum and distance to the center of the mass (m)
q1 — angle that the pendulum makes with the vertical
q2 — angle of the wheel
τ — input torque applied on the wheel (Nm)

The Lagrangian model of IWP possesses certain very interesting properties, for
instance, note that the inertial matrix M is constant as all the Christoffel Symbols
associated with M vanish. Thus IWP is a flat underactuated mechanical system
with kinetic symmetry.
The state space model of IWP is unsuitable for direct application of DSC,5 for
not being in the strict feedback form. Feedback Linearization7 employs a change
May 15, 2007 7:33 WSPC/123-JCSC 00351

84 N. Qaiser et al.

of control and coordinate transformation, which leaves the system dynamics linear
or at least partially linear, more amenable to control strategies like IBS, MSS, and
DSC. However, we use another tool, namely Collocated Partial Feedback Lineariza-
tion due to Spong,8 to circumvent the problem with the following change of control
and coordinate transformation9 :

τ = αu + β(q1 ) , (3)

where α = (m22 − m21 m12 /m11 ), β = (m21 w/m11 ) sin(q1 ), and

η1 = m11 q̇1 + m12 q̇2 ,


η2 = q1 , (4)
η3 = q̇2 .

This partially linearizes the system rendering IWP into strict feedback form

η̇1 = w sin(η2 ) , (5)


1 m12
η̇2 = η1 − η3 ,
m11 m11 (6)
η̇3 = u .

Note that q2 does not play any important role in dynamics of system; thus it is not
included as a state variable.

Remark 1. The system after coordinate transformation is a cascaded intercon-


nection of a linear subsystem, Eq. (6) and a nonlinear, core or reduced subsystem
Eq. (5), in the strict feedback form.

Remark 2. Spong et al.1 using a standard method from Isidori,7 performs exact
feedback linearization of the system. Afterwards IBS10 can be used for stabilization,
but it entails propagation of derivatives causing explosion of terms. Olfati-Saber10
applies further change of coordinates and control prior to proceeding for controller
design. Our design approach does not require any further transformations prior to
application of a very simple design algorithm.

3. Controller Design
The DSC technique is not applicable in usual fashion as the core system is non-
affine in control. Thus first assuming η2 as the virtual input and using the CLF
method, a stabilization function is found for the core subsystem. Afterwards, the
DSC technique is used to design a u forcing η2 to track the required stabilization
function, ultimately stabilizing the total system. DSC is chosen as it not only has
a nice trajectory tracking feature with an arbitrarily small bounded error but also
does not exhibit the phenomenon of explosion of terms associated with IBS.
May 15, 2007 7:33 WSPC/123-JCSC 00351

Exponential Stabilization of the IWP Using Dynamic Surface Control 85

3.1. Core subsystem controller design


Assuming η2 as an input for the core system and using the Control Lyapunov
method, a control α(η1 ) is found such that it stabilizes the origin of Eq. (5).
Taking V0 (η1 ) = 12 η12 as a Lyapunov function candidate
V̇0 (η1 ) = η1 η̇1 = wη1 sin(α(η1 )) . (7)
To render V̇0 negative definite we need α(η1 ) such that η1 α(η1 ) > 0 and
|α(η1 )| < π for all η1 . It is easy to verify that any sigmoidal function α(η1 ) sat-
isfies both conditions. We use α(η1 ) = − a tan−1 (c1 η1 ) as a virtual control input
that yields
V̇0 (η1 ) = − wη1 sin(a tan−1 (c1 η1 )) ≤ 0 ,
π (8)
0<a≤ , c1 > 0 ,
2
which implies Global Asymptotic Stability (GAS) for the origin of Eq. (5).

3.2. Outer subsystem controller design


To stabilize Eq. (5) the trajectory for η2 to follow is given as
η2d := α(η1 ) . (9)
Applying the DSC technique we design the desired control law. Theorem for
Boundedness of Tracking Error Using DSC for Lipschitz Systems by Swaroop et al.5
lists necessary requirements for DSC. It is trivial to verify that required assumptions
are satisfied by Eq. (5) with regard to the system and by Eq. (9) with regard to
the trajectory as
• w sin(η2 ) is C 1 ,
• η2d is twice differentiable and bounded with bounded derivatives up to the second
order.
Thus the design procedure can be applied easily. This however does not guar-
antee the stability of the closed-loop system, as the tracked trajectory is not
independent.

3.2.1. Design procedure


Let the error in generation of stabilization function (9) by η2 be S1
S1 := η2 − η2d , (10)
1 m12
Ṡ1 = η̇2 − η̇2d = η1 − η3 − η̇2d . (11)
m11 m11
Assuming η3 as next virtual input, it is chosen to derive S1 to zero; thus
 
m11 1
η̄3 = K 1 S1 + η1 − η̇2d (K1 > 0) . (12)
m12 m11
May 15, 2007 7:33 WSPC/123-JCSC 00351

86 N. Qaiser et al.

Derivative η̇2d is found by differentiating Eq. (9) as the derivation is trivial. A low
pass filter as demonstrated in the subsequent design procedure can be used even at
this stage to avoid derivative calculations, i.e., of η̇2d if it is complicated. However,
experience with simulations shows that use of a filter increases the control effort
substantially.
Define the second surface as
S2 := η3 − η3d , (13)
Ṡ2 = η̇3 − η̇3d = u − η̇3d . (14)
Control input u is designed to derive S2 to zero
u = η̇3d − K2 S2 (K2 > 0) . (15)
Notice that the direct calculation of η̇3d (t) required at this step by the conventional
backstepping design procedure leads to complexity due to “explosion of terms”.
Motivated by the DSC technique a low pass filter with small positive time constant
τ3 given below is used here
τ3 η̇3d (t) + η3d (t) = η̄3 , η3d (0) = η̄3 (0) . (16)
η3d (t) and η̇3d (t) are obtained by filtering η̄3
η̄3 − η3d (t)
η̇3d (t) = . (17)
τ3
Thus
η̄3 − η3d (t)
u= − K 2 S2 (18)
τ3
and
Ṡ2 = − K2 S2 . (19)
Substitution of u in Eq. (3) yields the required τ , completing the design. A com-
parison to existing designs presented in Refs. 1 and 10 reveals the ease of design
and the simplicity of the obtained control law.

3.3. Stability analysis


Core equation (5) with virtual control input given as Eq. (9) is GAS, and errors
are removed exponentially, heuristically, indicating the stability of the composite
system. In case of backstepping, stability of the closed loop follows from the GAS
of the core system. But in our case inclusion of the low pass filter requires special
attention. Proof of the following theorem justifies the heuristics.

Theorem 1. There exists ε∗ > 0 such that for all 0 < ε < ε∗ , the controller
designed in Eqs. (7)–(19) exponentially stabilizes the origin of systems (5) and (6)
for K1 , K2 , c1 > 0, ε = τ3 , and 0 < a ≤ π2 .
May 15, 2007 7:33 WSPC/123-JCSC 00351

Exponential Stabilization of the IWP Using Dynamic Surface Control 87

Proof. Let the boundary layer error be given as

z := η3d − η̄3 . (20)

Thus from Eqs. (20), (17), and (13)


z
η̇3d = − (21)
τ3
and

η3 = S2 + η̄3 + z . (22)

After substitutions and some algebraic manipulations


m12 m12
Ṡ1 = − K1 S1 − S2 − z. (23)
m11 m11
Boundary layer error dynamics can be found as
m12
z = η3d − η̄3 = η3d − K1 S1 + η̇2d , (24)
m11
 
z m11 m12 m12 η̇1
ż = − − K 1 − K 1 S1 − S2 − z − + η̈2d , (25)
τ3 m12 m11 m11 m12
where
ac21 sin(κ)[a cos(κ) + 2c1 η1 sin(κ)]
η̈2d =
(1 + c21 η12 )2
and

κ = S1 − a tan−1 (c1 η1 ) .

As filter time constant τ3 can be set arbitrarily small, the composite system
may be written as a Standard Singularly Perturbed Model given as follows.
ẋ1 = w sin(α(x1 ) + x2 ) ,
m12 m12
ẋ2 = − K1 x2 − x3 − z, (26)
m11 m11
ẋ3 = − K2 x3 ,

εż = g(x, z, ε) = − z3 + εp(x, z) , (27)

where x = [z1 , S1 , S2 ]T and ε = τ3 . Stability of the compositions (26) and (27)


implies that of IWP. To show the main result we use the following important
theorem11 for the stability of singularly perturbed models.

Theorem 2. Consider the singularly perturbed system

ẋ = f (t, x, z, ε) , (28)

εż = g(t, x, z, ε) . (29)


May 15, 2007 7:33 WSPC/123-JCSC 00351

88 N. Qaiser et al.

Assume that the following assumptions are satisfied for all (t, x, ε) ∈ [0, ∞) × Br ×
[0, ε0 ]:
(i) f (t, 0, 0, ε) = 0 and g(t, 0, 0, ε) = 0.
(ii) The equation 0 = g(t, x, z, 0) has an isolated root z = (t, x) such that h(t, 0) =
0.
(iii) Functions f, g, h and their partial derivatives up to the second order are
bounded for z − h(t, x) ∈ Bρ .
(iv) The origin of the reduced system ẋ = f (t, x, h(t, x), 0) is exponentially stable.
dy
(v) The origin of the boundary-layer system dτ = g(t, x, y + h(t, x), 0) is exponen-
tially stable, uniformly in (t, x).
Then there exists ε∗ > 0 such that for all ε < ε∗ , the origin of Eqs. (28) and (29)
is exponentially stable.
An isolated root, z := h(x) = 0, can be found as solution of 0 = g(x, z, 0),
and there is no need to shift the quasi-steady state of z to the origin; thus y :=
z − h(x) = z. It is trivial to verify that f (0, 0, ε) = 0 and g(0, 0, ε) = 0. Also the
functions f, g, h and their partial derivatives up to the second order are bounded
in balls given as z − h(x) ∈ Bρ and (x, ε) ∈ [0, ∞) × Br × [0, ε0 ].
The reduced system ẋ = f (x, h(t, x), 0) is given as follows:
ẋ1 = w sin(α(x1 ) + x2 ) ,
m12
ẋ2 = − K1 x2 − x3 , (30)
m11
ẋ3 = − K2 x3 .
For Eq. (30) it is trivial to verify that f (x) is continuously differentiable and its
Jacobian linearization
 
− ac1 w w 0
   
∂f   m12 
A= =  0 −K − 
∂x x=0 
1
m11 
0 0 − K2
is Hurwitz (all Re(λi ) < 0 in λi = {− K1 , − K2 , − ac1 w} for A), bounded, and
Lipschitz.11 By using the Lyapunov indirect method (Ref. 11, Theorem 4.15), these
properties imply exponential stability of the reduced system. Further, applying a
change of time scale to Eq. (27) as
dy dy dτ 1
ε = or =
dt dτ dt ε
yields the boundary layer system given as
dy
= g(x, y + h(t, x), 0) = − y ,

which is also exponentially stable. As filter time constant ε can be set arbitrarily
small, Eqs. (26) and (27) satisfy all the conditions listed in Theorem 2 for 0 < ε <
ε∗ , proving exponential stability of the closed-loop system.
May 15, 2007 7:33 WSPC/123-JCSC 00351

Exponential Stabilization of the IWP Using Dynamic Surface Control 89

4. Simulation Results
Initial condition effects on system stability and controller performance were stud-
ied numerically. For an objective performance comparison we use same system
parameters as used by Olfati-Saber10 and Spong,1 namely m11 = 4.83 × 10−3 ,
m12 = m21 = m22 = 32 × 10−6 , and w = 379.26 × 10−3 .
As obvious from the analysis, Ki can be set moderately high for faster conver-
gence rates. However, the actuation limits must be kept in mind during design as
higher values of Ki result in higher control effort too. The conventional DSC design
suggests to keep the outer loop faster, i.e., K2 > K1 , but as shown in stability
analysis and confirmed with simulations it is not necessary in this case. Filter time
constant τ3 controls boundary layer error, hence must be set as low as possible as
there is no lower limit. Again physical component values and actuator saturation
must be kept in mind as smaller values increase initial control effort peaks. It also
makes the control signal noisy and causes numerical problems in simulation as a
very small τ3 makes the system stiff. Average control effort and rise time can be
tuned by adjusting c1 . Following controller parameters were used for simulations:
a = π/2, c = 9, K1 = 4, K2 = 6, and τ3 = 0.035. In Figs. 2–5 the x-axis represents
time in seconds.
Initial position of the pendulum determining S1 appears as a disturbance dom-
inating the bounded stabilizing function α(z1 ). This seems destabilizing the core
system apparently. But for a variety of initial conditions as depicted in Fig. 2, sim-
ulations show these decay exponentially to arbitrarily bounded values, conforming
the heuristics. However, stability guarantee still comes from the state boundedness
property of the driven system.
As depicted in Fig. 3 the nonlinear controller stabilizes the pendulum from
its downward stable equilibrium point to its upright unstable equilibrium point.
The swing up is performed aggressively with negligible transients. The wheel stops
rotating in reasonable time as shown in Fig. 4.
The swing up and stabilization is faster than the design by Olfati-Saber10 and
Spong1 with less average control effort. However, initial control effort peaks are












     

Fig. 2. Convergence of error surface S1 for different initial conditions.


May 15, 2007 7:33 WSPC/123-JCSC 00351

90 N. Qaiser et al.

Pos (rad)−Vel (rad/s)


0

-5
angle
velocity
-10

0 1 2 3 4 5 6

Fig. 3. Pendulum position (rad) and velocity (rad/s).




 

     

Fig. 4. Wheel velocity (rad/s).




 




 
     

Fig. 5. Control effort (Nm).

comparatively larger (1.4 Nm versus 0.6 Nm by Olfati-Saber) in this case as shown


in Fig. 5. These peaks can be reduced by increasing filter time constant which of
course cannot cross an upper limit as indicated in stability analysis.
The controller does not need to switch to a local stabilizing controller in contrary
to what is done in passivity-based designs. Thus the system smoothly converge to
origin, refer to the phase portrait shown in Fig. 6. The design involves cancellation
of undesired terms in Eq. (11) involving m11 and m12 . When the parameter iden-
tification errors exist, this can affect the accuracy and stability of the method. To
check these aspects simulations were also done with parametric errors too.
May 15, 2007 7:33 WSPC/123-JCSC 00351

Exponential Stabilization of the IWP Using Dynamic Surface Control 91


    





     

 
 

Fig. 6. State trajectory in the (q, q̇) plane.

Although the controller was designed without robustification in mind, it was


able to stabilize the system with variations up to ± 20% in system parameters
around nominal values at which the controller was designed. However, some effects
on transient behavior of the system were observed.
Simulation results are satisfactory keeping in mind especially the design ease
and control law simplicity.

5. Conclusions
The DSC technique has been exploited in a novel way to design a new controller for
the nonlinear IWP system. The model is brought to the strict feedback form before
applying the DSC. Stability of the system is analyzed using concepts from singular
perturbation theory. Some critical issues concerning the initial conditions of the
system and robustness are studied numerically. Design simplicity is demonstrated,
and controller performance is compared to existing designs using both theoretical
and simulation studies. In conclusion, the proposed controller has a relatively simple
design procedure requiring fewer transformations and resulting in a less complicated
control law, which achieves faster stabilization. The structure is also simpler requir-
ing no supervisory switching controller. Further improvements envisioned include
determination of maximum value of ε, achieving the robustness and generalization
of the scheme to cover the whole subclass of nonlinear underactuated systems.

Acknowledgment
The authors like to acknowledge Higher Education Commission of Pakistan for
funding this research.
May 15, 2007 7:33 WSPC/123-JCSC 00351

92 N. Qaiser et al.

References
1. M. W. Spong, P. Corke and R. Lozano, Nonlinear control of the Inertia Wheel Pen-
dulum, Automatica 37 (2001) 1845–1851.
2. I. Fantoni, R. Lozano and M. Spong, Passivity based control of the pendubot, Proc.
American Control Conf., San Diego, CA, June 1999, pp. 268–272.
3. E. D. Sontag and H. J. Sussman, Further comments on the stability of the angular
velocity of a rigid body, Syst. Control Lett. 12 (1988) 213–217.
4. J. K. Hedrick and Y. Y. Yip, Multiple sliding surface control: Theory and application,
Trans. ASME, J. Dyn. Syst. Meas. Control 122 (2000) 586–593.
5. D. Swaroop, J. K. Hedrick, P. P. Yip and J. C. Gerdes, Dynamic surface control for
a class of nonlinear systems, IEEE Trans. Automat. Control 45 (2000) 1893–1899.
6. H. Goldstein, Classical Mechanics, 2nd edn. (Addison-Wesley, USA, 1981).
7. A. Isidori, Nonlinear Control Systems, 2nd edn. (Springer-Verlag, Berlin, New York,
1989), Chapter 4.
8. M. W. Spong, Energy based control of a class of under actuated mechanical systems,
IFAC World Congress, July 1996.
9. R. Olfati-Saber, Normal forms for underactuated mechanical systems with symmetry,
IEEE Trans. Automat Control 47 (2002) 305–308.
10. R. Olfati-Saber, Nonlinear control of underactuated mechanical systems with applica-
tion to robotics and aerospace vehicles, Ph.D. thesis, Department of Electrical Engi-
neering and Computer Science, MIT, USA (2001).
11. H. K. Khalil, Nonlinear Systems, 3rd edn. (Prentice-Hall, USA, 2000), pp. 165, 423–
435, 589.

View publication stats

You might also like