You are on page 1of 7

IEEE TRANSACTIONS ON ROBOTICS AND AUTOMATION, VOL. 20, NO.

2, APRIL 2004

359

Swing-Up and Stabilization of an Underactuated


Manipulator Without State Feedback of Free Joint
Hiroshi Yabuno, Kazukuni Goto, and Nobuharu Aoshima
AbstractA technique without state feedback of the free link is theoretically and experimentally proposed for the swing-up and the stabilization
at the upright position of the free link in a two-link underactuated manipulator. An appropriate actuation of the bifurcations produced in the free
link realizes the control objectives.
Index TermsBifurcation, dynamic stabilization, high-frequency excitation, pitch-fork bifurcation, underactuated manipulator.

I. INTRODUCTION
Control of underactuated manipulators has received increasing attention in recent years. The manipulators with free or passive joints
(links) are called underactuated manipulators and in these systems, the
number of generalized coordinates is larger than the number of control inputs (the number of actuators). Also, the equations of motion for
the free or passive links are homogeneous, and they can be regarded as
dynamic constraints. The constraints are determined as nonholonomic
constraints because they are nonintegrable. From the practical point of
view, investigation into the motion control of the underactuated manipulator is very interesting for the realization of the reduction of the
weight, energy consumption, and cost of the manipulator, and for overcoming actuator failure due to unexpected accidents. There have been
many studies on the control of underactuated manipulators (comprehensive references can be found in [1] and [2]). Arai and Tachi [3]
theoretically and experimentally proposed a method of controlling the
position of a two-link underactuated manipulator with a brake at the
passive joint by using the coupling characteristics of manipulator dynamics. Yu et al. [4] presented a position control method for an underactuated manipulator without a brake which used the friction at the
free joint. Nakamura et al. [5] theoretically and experimentally discussed the positioning of both free and active joints of two-link underactuated manipulators by means of periodic inputs to the active joint.
A energy-based control method for swing-up and stabilization is also
proposed for a model equivalent to the system in the present paper [6].
In the above-discussed methods, the control of the free link is carried
out by actuating the active joint based on the motion of the free link.
Hence, the information related to the motion of the free link, i.e., the
measurement of the angle of the free joint, is essential.
In this paper, a control method for a two-link underactuated manipulator under the gravity effect, in which the second joint (free joint)
lacks not only an actuator but also a sensor, is proposed (the proposed
method can also be regarded as a control strategy for the case when
not only the actuator but also the sensor breaks down). The control objective is to swing up the second (free) link, which hangs down in the
gravity direction in the initial state, to the state where the second (free)
link points in the direction opposite the gravity effect (hereafter, we call
Manuscript received November 15, 2002; revised March 29, 2003. This paper
was recommended for publication by Associate Editor H. Arai and Editor A. De
Luca upon evaluation of the reviewers comments. This work was supported by
the Japanese Ministry of Education, Culture, Sports, Science and Technology,
under Grants-in-Aid for Scientific Research 13650243 and 09650454.
H. Yabuno and N. Aoshima are with the Institute of Engineering Mechanics
and Systems, University of Tsukuba, Tsukuba 305-8573, Japan (e-mail:
yabuno@esys.tsukuba.ac.jp; aoshima@esys.tsukuba.ac.jp).
K. Goto is with Fuji Heavy Industry Ltd., Ohizumi 370-0514, Japan (e-mail:
gto_wa@zdb.so-net.ne.jp).
Digital Object Identifier 10.1109/TRA.2004.824692

Fig. 1. Analytical model.

this state the upright position), and also to stabilize the free link at this
state. The swing-up and stabilization are accomplished without feedback with respect to the angle and angular velocity of the free joint.
The dynamic stabilization phenomenon under high-frequency excitation [7][10], which is also known as a method of stabilizing the inverted pendulum without feedback control, is introduced. Recently, the
application of the dynamic stabilization phenomenon to the underactuated manipulator has also been theoretically proposed [11]. However,
in these studies on the dynamic stabilization and its application, there
has been no mention of a strategy for controlling the motion of the
free link from one position to the unstable equilibrium position, i.e.,
swing-up in the present paper, or of the condition with respect to the
excitation amplitude and frequency for realization of the swing-up.
In the present study, an unified strategy for swing-up and stabilization at the upright position is theoretically proposed without state
feedback control of the second joint, by using the nonlinear characteristics [12] of the bifurcation produced in the second link under the
high-frequency excitation. Also, the condition for the realization of
the swing-up and the stabilization is analytically shown according to
bifurcation theory [13]. By using the actuator attached to the first joint,
the first link can have any configuration with respect to the direction of
the gravity effect, and can also be excited with high frequency around
the configuration (hereafter, the angles of the first link expressing the
configuration with respect to the direction of the gravity effect are
called offset of the excitation). Then it is theoretically clarified that
depending on the offset of the excitation, the second link undergoes
supercritical and subcritical pitchfork bifurcations [13] and their perturbed bifurcations [14]. In these bifurcations, there are different stable
equilibrium states. The swing-up of the second link is accomplished
without state feedback control of the second joint by the variation of
these equilibrium states under an appropriate change of the offset of
the excitation with time, i.e., of the perturbations of the bifurcations in
the second link. After the accomplishment of swing-up to the upright
position, the second link is stable without state feedback control of
the second joint, because at this position, the second link undergoes
the subcritical pitchfork bifurcation and has the local minimum of the
potential energy at the upright position. Finally, the validity of the
proposed control method is experimentally confirmed.
II. TWO-LINK UNDERACTUATED MANIPULATOR
A. Analytical Model of an Underactuated Manipulator
The analytical model of an underactuated manipulator is shown in
Fig. 1. The manipulator is moving on the plane  0 which is inclined at
angle from the horizontal plane  . Then, the manipulator experiences
the gravity effect g 0 (= g sin ) [m=s2 ] in the positive direction of the
x0 axis. The system is controllable in the neighborhood of the upright
position in linear sense as long as g sin is not zero, in contrast with

1042-296X/04$20.00 2004 IEEE

360

IEEE TRANSACTIONS ON ROBOTICS AND AUTOMATION, VOL. 20, NO. 2, APRIL 2004

the studies of Arai et al. [1], Hong [2], Arai and Tachi [3], and Yu et
al. [4].
Parameters, mi , li , and lig in Fig. 1 show the ith mass, length of the
ith link, and the distance between the ith joint and the center of gravity
of the ith link, respectively. Parameters Ii is a mass moment of inertia
about the center of the ith link. The parameter values corresponding to
the subsequent experimental apparatus are as follows:

m1 = 1:50 2 1001 kg; m2 = 7:18 2 1002 kg


l1 = 1:23 2 1001 m; l2 = 0:19 m
l1g = 0:70 2 1001 m; l2g = 0:55 2 1001 m
I1 = 8:71 2 1004 kg 1 m2 ; I2 = 5:49 2 1004 kg 1 m2
g = 7:92 2 1001 m/s2 ( = 8:10 2 1002 rad):

1 = a 

The first joint has an actuator which can give torque  for the first link
and control the position of the first link easily. On the other hand, the
second joint is a free joint because it has no actuator. Therefore, in the
static equilibrium state where the first link is not actuated, the second
(free) link points to the positive direction of the x0 axis, regardless of
the configuration of the first active link. Our control objective for such
an underactuated manipulator is to swing up the second link from the
positive direction of the x0 axis to the negative direction of the x0 axis
(upright position) and to stabilize the second link at this position.
B. Equation Governing the Motion of the Second Link
The total kinetic energy of the first and the second link T and potential energy U are shown as follows:

1
2

2 2
(I1 + m1 l1g )_1

1
2
1

2 + _2 + (1 + c cos 2 )(0a1 cos t3 + 1o )


2
2 3
3
2
+ c sin 2 (a1 sin t 0 2a1 sin t 1 _1o + _1o )
3
+  sin(a1 cos t + 1o + 2 ) = 0

(1)

dt

A2

d2 

2
+ (A2 + 1 cos 2 )
dt2

d2 

+ 3 sin(1 + 2 ) = 0;

dt

1
+ 1 sin 2
dt2

(7)

where (1) denotes the derivative with respect to dimensionless time,


= t 1 ! , and the dimensionless parameters, c and  , are expressed
as follows:

t3

(2)

The equations of motion can be expressed by using Lagranges formulation as follows:


d2 1
(A1 + A2 + 2 1 cos 2 )
dt2
2
d1 + d2 d2 sin 
d
2
+ (A2 + 1 cos 2 )
2
2 0 1 2
+ 2 sin 1 + 3 sin(1 + 2 ) =

(5)

Next we rewrite the above equation in the dimensionless form. Using


the inverse value of the frequency ! as the representative time, the dimensionless equation governing the motion of the second link is obtained from (6), as follows:

cos !t + 1o

where the first term is for giving high-frequency excitation to the


second link (the frequency ! is called the excitation frequency),
and the second term expresses the configuration of the first link,
with respect to the direction of the gravity effect as the previously
mentioned offset of the excitation for producing the perturbations
in the bifurcations under the high-frequency excitation, as detailed in
Section III. Substituting (5) into (4) yields the equation governing the
motion of the second link as follows:
2
d2 2 + 1 + 1 cos 
0!2 a1 cos !t + d dt1o
2
2
dt2
A2
2
1 sin  0!a sin !t + d1o
+
2
1
A2
dt

3
+
sin(a1 cos !t + 1o + 2 ) = 0:
(6)
A2

(I2 + m2 l1 + m2 l2g + 2m2 l1 l2g cos 2 )_1

I2 + m2 l22g _22
2
+ I2 + m2 l2g + m2 l1 l2g cos 2 _1 _2
0
U = 0 m1 g l1g cos 1 0 m2 g0
2 [l1 cos 1 + l2g cos(1 + 2 )] :
+

motion control of the second link. The motion of the first link is influenced by the motion of the second link, as shown in (3). However, it is
assumed here that the position control of the first link can be approximately realized by feedback control only with respect to the angle and
angular velocity of the first link. Hence, we assume that the motion of
the first link is not affected by the motion of the second link, and then
we can ignore (3). Now we can set the position of the first link as follows:

dt

dt

d1
dt

(3)

(4)

where the constant parameters, Ai and i , are expressed as follows:

A1 = I1 + m1 l12g + m2 l12 (2:69 2 1003 rad)


A2 = I2 + m2 l22g (7:66 2 1004 kg m2 )
1 = m2 l1 l2g (4:86 2 1004 kg m2 )
2 = (m1 l1g + m2 l1 )g0 (1:53 2 1002 kgm2 =s2
3 = m2 l2g g0 = m2 l2g g sin (3:13 2 1003 kgm2 =s2 )
where the values in parentheses correspond to those of the subsequent
experimental apparatus.
In this research, we excite the angle of the first link 1 sinusoidally,
and then by using the motion of the tip of the first link, we carry out the

c=

1 ;  = A :
A2
!2

(8)

In the above equation, a viscous damping effect _2 is also considered


as the friction existing inherently and necessarily in general purpose
radial bearings. The parameter values corresponding to the subsequent
experiment are as follows:

c = 0:635  = 4:08
 = 0:524
(9)
!2
!
where ! is variable and  is experimentally identified under the assumption of logarithmic decrement [15].
III. NONLINEAR ANALYSIS FOR THE BEHAVIOR
UNDERACTUATED MANIPULATOR

OF THE

A. Simplification of the Governing Equation by Nonlinear Analysis


Because we realize the control objective by using the dynamic stabilization phenomenon, we set the magnitude of the excitation frequency
! to be a large value, compared with the linear natural frequency of
the second pendulum 3 =A2 , and then the dimensionless parameter
 is very small. Also, we set a small value for the excitation amplitude a1 . Now to perform nonlinear analysis, we scale the parameters,
,  , and a1 , according to the general procedure for systems under
high-frequency excitation [16], so that at the zeroth-order O(0 ) in the

IEEE TRANSACTIONS ON ROBOTICS AND AUTOMATION, VOL. 20, NO. 2, APRIL 2004

perturbation procedure, the influence of the inertia 2 does not balance


any terms, and at the first-order O  , the influence of the damping 2
3
t , as shown later [(14) and
balances the influence of excitation a1
as a book(15)]. Hence, by using a small parameter  <  
keeping device, we quantitatively set the magnitudes of the parameters
as follows:

()

cos

(0

361

Furthermore, substituting (18) and (20) into (16) yields


2

D0 22

1)

= 2^a1 cD1 C0 sin C0 sin t0 0 D12 C0


0 ^ a^1 (1 + c cos C0)sin t0 + D1 C0
+ 21 a^21 c c cos C0 sin C0(1 + cos2t0 )

= 2 ^ a1 = a^1  = ^ (^ = a^1 = ^ = O(1)) (10)


where (^) denotes of the order O(1). Then the dimensionless equation


+ 2sin C0 cos 2t0

is written as follows:

 + ^_2 + (1 + c cos 2 )(0a^1 cos t3 + 1o )


2 )
+ c sin 2 (2a^21 sin2 t3 0 2a^1 sin t3 1 _1o + _1o
2
3
(11)
+  ^ sin(a^1 cos t + 1o + 2 ) = 0:

2

We analyze (11) using the method of multiple scales [17], which is


a kind of perturbation method which has conventionally been used for
solving nonlinear differential equations. We introduce the multiple time
scales as follows:

= t3
3
2 3
3 3
t1 = t ; t2 =  t ; t3 =  t ; . . .

t0

(12)

where t0 is the fast time scale, and t1 , t2 , and t3 are slow time scales
[17]. We seek an approximate solution in the form
2

= 20 + 21 + 2 22 1 1 1:

(13)

Substituting (13) into (11) and equating the coefficients of like powers
of  yield the following equations for the orders:

( 0 ): D0220 = 0
(14)
1
2
O ( ): D0 21 = 02D0 D1 20 0 
^D0 20
+ a^1 c cos t0 cos 20 + a^1 cos t0 (15)
2
2
2
O ( ): D0 22 = 02D0 D1 21 0 (2D0 D2 + D1 )20
0 ^(D0 21 + D1 20 )
0 a^21 c sin2 t0 sin 20
0 a^1 c21 sin 20 cos t0
0 ^ sin(1o + 20 )
(16)
where Di = @=@ti . We assume that the change of 1o is so slow
that the effect is detected by the slow time scale t3 . Then _1o (t3 )
and 1o (t3 ) are expressed as _1o (t3 ) = 3 D3 1o and 1o (t3 ) =
6 2
O 

 D3 1o , respectively, and no terms concerned with the time derivative of 1o appear in (14)(16), which are considered until the accuracy of O 3 .
From (14), we obtain 20 as follows:

( )

20

= C 1 t0 + C 0 (t1 ; t 2 ; t 3 )

(17)

where C0 and C1 are integral constants. We note that the first term is
a secular term [17]. For a uniform expansion, this term must be eliminated by setting C1 to zero. Then the general solution becomes
20

= C 0 (t1 ; t 2 ; t 3 ):

= a^1 (1 + c cos C0)cos t0 :

(19)

The right-hand side does not contain any terms that produce secular
terms in 21 . The particular solution of (19) is
21

= 0a^1 (1 + c cos C0 )cos t0 :

(21)

Because the terms which do not explicitly contain t0 produce a secular


term in 22 , the sum of these terms must be set to zero, as follows:
2

D1 C0

+ ^D1 C0 + ^ sin(1o + C0)

0 c a2^1 sin C0 cos C0 = 0:


2 2

(20)

(22)

Then, multiplying both sides by 2 yields the following equation:


2 2
c a1
 C0
C0

1o
:
(23)
C0 0
C0
C0

 + _ + sin(

2 sin cos = 0
Because from (13) and (18) 2 is equal to C0 (= 20 ) in neglecting the
error of O(), we can approximately replace C0 in (23) by 2 . There+ )

fore, the equation governing the motion of the second link can be approximately described as follows:
2 2
c a1
2
 2
2

1o
:
(24)
2
2 0

 + _ + sin(

2 sin cos = 0

+ )

As a result, the governing equation (11), which is nonautonomous, is


transformed into the autonomous differential equation (24) by the nonlinear analysis using the method of multiple scales, and then it is very
easy to perform bifurcation analysis.
B. Bifurcation Analysis
Equation (24) with d2 2 =dt2
equation as follows:

f ; 1o ; 2eq

= d2=dt = 0 leads to the bifurcation


2 2

) : =  sin(1o + 2eq ) 0 c a21 sin 2eq cos 2eq


= 0:
(25)

The solutions 2eq express the equilibrium states of the second link. To
determine the stabilities of the equilibrium states, we superimpose on
them a small disturbance 2 , and obtain

1
2 (t) = 2eq + 12 (t):

(26)

Substituting (26) into (24) yields

12 + 1_2 +

2 2

0 c 2a1 (cos 2eq 0 sin 2eq ) 2 12 = 0:


(27)

(18)

We substitute (18) into (15) and obtain


D0 21

0 ^ sin(1o + C0 ):

By calculating the characteristic exponent of the above equation for


each equilibrium state, the stabilities of the equilibrium states are ex, 1o = ,
amined. In this section, for the special cases of 1o
= , we show the variation of the equilibrium states, deand 1o
pending on the dimensionless parameter expressing the magnitude of
the excitation frequency of the first link  .
Fig. 2 is the bifurcation diagram which shows the relationship between the excitation frequency of the first link and the equilibrium

= 4

=0

= 2

362

IEEE TRANSACTIONS ON ROBOTICS AND AUTOMATION, VOL. 20, NO. 2, APRIL 2004

Fig. 4. Bifurcation diagram of


Fig. 2. Bifurcation diagram of

Fig. 3.

Bifurcation diagram of

4).

= 0).

2).

states of the second link in the case of 1o = 0, where the solid and
dashed lines denote the stable and unstable equilibrium states, respectively. In the case of (= ( 3 =A2 )=! 2 ) > c2 a21 =2, i.e., when the
excitation frequency ! is not sufficiently high, the second link only
has the stable equilibrium state 2eq = 0, where the second link hangs
down in the direction of the gravity effect x0 . However, the second link
undergoes a supercritical pitchfork bifurcation at  = c2 a21 =2, and
then the equilibrium state 2eq = 0 is changed to an unstable equilibrium state. Also, as shown in Fig. 2, the stable nontrivial equilibrium
states bifurcate from the bifurcation point  = c2 a21 =2. Therefore,
in the case of  < c2 a21 =2, i.e., when the excitation frequency !
is sufficiently high, the second link has stable nontrivial equilibrium
states (2eq 6= 0), and is swung up from the positive direction of the
x0 axis. Furthermore, the angle 2eq of the stable equilibrium state becomes larger as the excitation frequency ! becomes higher ( becomes
lower).
Next, we consider the behavior of the second link in the case of
1o = =2. Fig. 3 is the bifurcation diagram which shows, in the case
of 1o = =2, the relationship between the excitation frequency of
the first link and the equilibrium states of the second link. In the case of
 > c2 a21 =2, i.e., when the excitation frequency ! is not sufficiently
high, the second link has only one stable equilibrium 2eq = 0=2,
where the second link hangs down in the direction of the gravity effect,
i.e., the positive direction of the x0 axis (absolute angle of the second
link is 2abs = 1 + 2eq = =2 0 =2 = 0). However, the second
link undergoes a subcritical pitchfork bifurcation at  = c2 a21 =2,
and then another equilibrium state, 2eq = =2, where the absolute
angle of the second link is 2abs = 1 + 2eq = =2 + =2 = 
(upright position), is changed to be stable. Also, as shown in Fig. 3,

the unstable nontrivial equilibrium states bifurcate from the bifurcation point,  = c2 a21 =2. Therefore, the upright position of the second
link can be stabilized without state feedback of the joint, in the case
of  < c2 a21 =2, i.e., when the excitation frequency ! is sufficiently
high. Then the upright position is stable against the disturbance that
is the angle in the hatched region bounded by the unstable nontrivial
equilibrium states, in the case of small angular velocity disturbance,
because the potential energy has the local minimum and maximum at
the upright position (stable equilibrium state) and the unstable equilibrium states, expressed with the dotted line in Fig. 3 [18], respectively.
Finally, we consider the behavior of the second link in the case of
1o = =4, which corresponds to the case when the supporting point
of the second link, i.e., the second joint, is excited in the direction
which is not parallel or perpendicular to the direction of the gravity
effect. Fig. 4, obtained from (25) to (27), includes simultaneously the
perturbed supercritical and perturbed pitchfork bifurcations, which are
found, for example, in the compressive buckled two-link system with
initial imperfection [14] and in an electromagnetic buckled two-link
system with initial imperfection [19], respectively. The combination of
the equilibrium states, (a)0 , (b)0 , and (c)0 , is regarded as the bifurcation
diagram of the perturbed supercritical pitchfork bifurcation [14], and
the combination of the equilibrium states, (b)0 , (c)0 , and (d)0 , can be
regarded as the bifurcation diagram of the perturbed subcritical pitchfork bifurcation [19].

IV. PROPOSITION OF A CONTROL METHOD


In this section, we theoretically propose a control method to swing
up the second link from the direction of the gravity effect (the positive
direction of x0 ) to the upright position (the negative direction of x0 ),
and to stabilize it at the upright position, by using the bifurcation phenomena and their perturbations shown in Section III. In the bifurcation
diagram of Fig. 3 (1o = =2), we focus on the case when the excitation frequency is sufficiently high ( < c2 a21 =2). The upright position
(2abs = 1 + 2eq  1o + 2eq =  ) is a stable equilibrium state.
However, since the other state in which the second link hangs down
(2abs = 1 + 2eq  1o + 2eq = 0), i.e., the initial position, is
also a stable equilibrium state, the swing-up from 2abs = 0 cannot
be realized by the excitation of the first link with 1o = 0. Therefore, we swing up the second link by using the variation of the stable
equilibrium state, depending on the value of 1o , which is shown in
Figs. 24. We show the continuous variation of the equilibrium states
by the change of the value of 1o in Fig. 5, which is the so-called
equilibrium surface [20] in bifurcation theory. The figure includes all
the bifurcation diagrams shown in the previous section; the branches
indicated by alphabets with no prime, single prime, and double prime

IEEE TRANSACTIONS ON ROBOTICS AND AUTOMATION, VOL. 20, NO. 2, APRIL 2004

Fig. 5.

Equilibrium space.

correspond to the branches of the supercritical, subcritical, and perturbed pitchfork bifurcations shown in Figs. 24, respectively. From
Fig. 5, we notice that the upper branch (c) of the bifurcation diagram in
Fig. 2, which expresses the stable equilibrium state, continuously leads
to the trivial stable equilibrium state (c)00 in Fig. 3 through stable equilibrium state (c)0 in Fig. 4, with increasing the value of 1o . Therefore,
the swing-up of the second link can be realized by changing the 1o
of the first link under the constant excitation frequency ! , i.e., under
the constant  .
For example, we can swing up the second link according to arrow on
the plane of  = a in Fig. 5. First we excite the first link with  = a
(constant excitation frequency ! = !a ) and 1o = 0, according to
(5). Due to the supercritical pitchfork bifurcation, the second link points
to the position on the upper branch (c) after the transient state. Then we
change the value of 1o from 0 to =2. As a result, the second link can
be swung up to the upright position (the absolute angle of the second
link: 2abs = 1o + 2eq = =2 + =2 =  ). Furthermore, the final
position of the second link is stable for the following reason. The subcritical pitchfork bifurcation occurs as shown in Fig. 3, corresponding
to the branches (c)00 and (d)00 in Fig. 5. Therefore, in the case of small
angular velocity disturbance, stability is maintained against the angular
disturbance, which is the angle in the hatched region bounded by the
unstable nontrivial equilibrium states, shown in Fig. 3.
We briefly comment on the case when the initial angle of 2 is
slightly negative at 1o = 0. In this case, the second link may point
to the lower branch of the supercritical pitchfork bifurcation in Fig. 2.
However, we can avoid such behavior by initially setting the offset
1o to be a small negative value and starting the excitation. Then, the
second link must be attracted to the upper branch of the supercritical
pitchfork bifurcation perturbed by the small negative offset, because
the second link points to the x0 direction, and the relative angle
2 is positive. Furthermore, because this upper branch is smoothly
connected to the upper branch of the supercritical pitchfork bifurcation
in Fig. 2, the second link must point to the point on the upper branch
in Fig. 2 when the offset is changed from the negative value to 0.
Next, let us investigate the method of determining the suitable excitation frequency and amplitude, i.e.,  and a1 . We show the variation of
the equilibrium states under changing the value of 1o in Figs. 6 and 7
for the cases of the sufficiently high constant excitation frequencies of
the first link ( = 8:82 2 1005 : = a ), and not sufficiently high excitation frequency ( = 9:97 2 1005 := b ), respectively. These figures
correspond to the intersections of the planes  = a and  = b with
the equilibrium space of Fig. 5, and also schematically show the configurations of the manipulator for each value of 1o . In the former case
of  = a , the swing-up can be realized, and then the upright position

363

Fig. 6. Bifurcation diagram (

4).

Fig. 7. Bifurcation diagram (

4).

Fig. 8.

Bifurcation set.

is stable. On the other hand, in the case of  = b , the swing-up cannot


be realized, because the stable equilibrium states are not continuously
varied with the change of 1o . Hence, in order to realize the swing-up,
it is necessary that there are four equilibrium states, including unstable
ones, for all values of 1o between 0 and =2. Next, we seek the parameter region for the above condition. In the left region of the line in
Fig. 5, which connects the bifurcation points for each value of 1o ,
there are four equilibrium states. The line, which corresponds to the
so-called bifurcation set in bifurcation theory, is theoretically obtained
from (25) and @f =@2eq = 0 and is shown on the plane  0 1o as
Fig. 8. In the hatched region, there are four equilibrium states, and in
the nonhatched region, there are only two equilibrium states in the section between 1o = 0 and 1o = =2. The arrow denotes the change

364

Fig. 9.

IEEE TRANSACTIONS ON ROBOTICS AND AUTOMATION, VOL. 20, NO. 2, APRIL 2004

Underactuated manipulator.

of 1o . In the case of  = b where the swing-up cannot be realized,


it can be seen that the arrow passes the nonhatched region. Finally, we
conclude from Fig. 8 that the setting of the sufficiently high-frequency
excitation ( < c2 a21 =4 := cr ) realizes the swing-up of the second
link and the stabilization at the upright position. In the parameters for
the subsequent experiment, we have cr = 9:07 2 1005 .
V. EXPERIMENT
We experimentally investigate the validity of the theoretically proposed control method by using the experimental apparatus shown in
Fig. 9. The first link is actuated by an AC servo motor with a rotary encoder [Mitsubishi Corp., HC-KFS43 (maximum torque: 3.8 Nm)]. The
second link has only an encoder [Koyo Corp., TRD-SH2500A (2500
pulse/rotation)], which is not used for the control of the second link,
but for collecting experimental data. The values of the parameters of
the experimental setup are shown in Section II-A.
We show in Fig. 10 the experimental time histories of the first and
second links for the case of ! = 213 rad/s (34 Hz), ( = a =
05
8:82 2 1005 < cr = 9:07 2 10 ), and a1 = 0:03 rad. First, we
set the offset of the excitation of the first link as 1o = 0 for 50 s,
and sinusoidally excite the first link. Then the second link is swung
up to 2  0:7 rad through the supercritical pitchfork bifurcation.
After that, we change the value of 1o until =2 rad with the speed
of d1o =dt = 0:1 rad/s. When the first link reaches =2, the relative
angle of the second link first reaches 2 = =2, and the swing-up of
the second link is accomplished (the absolute angle of the second link
is 2abs = 1o + 2 =  ) without state feedback control, with respect
to the angle and angular velocity of the second joint (2 and d2 =dt).
Finally, we keep the value of 1o at =2, and then the state of the
second link is also stabilized without state feedback of the second joint.
By the way, the speed of the change of 1o , d1o =dt, is determined
such that _1o is the order of O(3 ), according to the assumption used
in the theoretical analysis. In this experiment, the dimensionless speed
of the change is _1o = (d1o =dt)(1=!) = 0:1=213  5 2 1004 .
This value is the same order as the magnitude of  = 8:82 2 1005
( is regarded as O(2 ) in the analysis), and is larger than the order
of O(3 ) assumed in the analysis. Hence, the speed of the change set
in this experiment is a little faster than the speed assumed in the theoretical analysis. However, as a result, even in such a hard condition,
the swing-up to the upright position is carried out because the speed is
not so fast that the transient state does not excessively deviate from the
equilibrium states on the equilibrium surface in Fig. 5.

Fig. 10. Motion of the underactuated manipulator (


rad/s. (a) Time history of . (b) Time history of ).

= 0 03,

= 213

Next, we perform the experiment for ! = 201 rad/s (32 Hz:  =


2 1005 ). The methods of changing the value of 1o and
other excitation parameters are the same as those in the above experiment. In this case, the swing-up cannot be realized, because the excitation frequency is not sufficiently high. Hence, in this case, the condition necessary to accomplish the swing-up ( < cr ) is not satisfied
05 > cr = 9:07 2 1005 ), and while 1o is
( = a = 9:97 2 10
changed from 0 to =2, the situation where only two equilibrium states
exist occurs as shown in Fig. 7. This result corresponds to that in Fig. 8,
where the arrow indicating the change of 1o passes through the white
region as arrow (i).
b = 9:97

VI. CONCLUSIONS
This paper addresses the control of a two-link underactuated manipulator (the first link is directly actuated, and the joint connecting the
first and second links does not have an actuator or a sensor) under the
gravity effect. Utilizing the dynamics stabilization phenomenon under
a sinusoidally high-frequency excitation of the first link, the second
(free) link is swung up from the gravity direction to the opposite direction (upright position) and is stabilized at this position, under the
condition of no information, with respect to the motion of the second
link (without state feedback of the free joint). From bifurcation analysis for the autonomous equation obtained by the method of multiple
scales, it is analytically shown that changing the configuration of the
first link with respect to the gravity direction, i.e., the offset of the excitation, causes the variation of the stable equilibrium states of the second
link, because the perturbations of the bifurcation due to the high-frequency excitation are varied, depending on the offset of the excitation.
As a result, it is theoretically clarified that the control objective can be
realized without state feedback of the second joint. Furthermore, experiments using a simple apparatus are performed. The validity of the
theoretically proposed control method is experimentally confirmed.

IEEE TRANSACTIONS ON ROBOTICS AND AUTOMATION, VOL. 20, NO. 2, APRIL 2004

365

ACKNOWLEDGMENT

Stable Teleoperation With Time-Domain Passivity Control

The authors express their thanks to Mr. T. Matsuda, a graduate student of University of Tsukuba, for his assistance.

Jee-Hwan Ryu, Dong-Soo Kwon, and Blake Hannaford

REFERENCES
[1] H. Arai, K. Tanie, and N. Shiroma, Nonholonomic control of a
three-DOF planar underactuated manipulator, IEEE Trans. Robot.
Automat., vol. 14, pp. 681695, Oct. 1998.
[2] K. S. Hong, An open-loop control for underactuated manipulators
using oscillatory inputs: Steering capability of an unactuated joint,
IEEE Trans. Contr. Syst. Technol., vol. 10, pp. 469480, May 2002.
[3] H. Arai and S. Tachi, Position control of a manipulator with passive
joints using dynamic coupling, IEEE Trans. Robot. Automat., vol. 7,
pp. 528534, Aug. 1991.
[4] K. H. Yu, Y. Shito, and H. Inooka, Position control of an underactuated
manipulator using joint friction, Int. J. Non-Linear Mech., vol. 33, pp.
607614, 1998.
[5] Y. Nakamura, T. Suzuki, and M. Koinuma, Nonlinear behavior and control of a nonholonomic free-joint manipulator, IEEE Trans. Robot. Automat., vol. 13, pp. 853862, Dec. 1997.
[6] I. Fantoni, L. Lozano, and M. W. Spong, Energy-based control of the
pendubot, IEEE Trans. Automat. Contr., vol. 45, pp. 725729, Apr.
2000.
[7] P. L. Kapitza, Dynamical stability of a pendulum when its point of suspension vibrates, and pendulum with a vibrating suspension, in Collected Papers of P. L. Kapitza, D. T. Haar, Ed. London, U.K.: Pergamon, 1965, vol. 2, pp. 714737.
[8] A. Stephenson, On a new type of dynamical stability, Mem. Proc.
Manch. Lit. Phil. Sci., vol. 52, pp. 110, 1908.
[9] R. E. Bellman, J. Bentsman, and S. M. Meerkov, Vibrational control
of a class of nonlinear systems: Vibrational stabilization, IEEE Trans.
Automat. Contr., vol. AC-31, pp. 710716, Aug. 1986.
[10] S. P. Weibel and J. Bailliwul, Open-loop oscillatory stabilization of an
-pendulum, Int. J. Control, vol. 71, pp. 931857, 1998.
[11] T. Suzuki and Y. Nakayama, Anti-gravity control of free-joint manipulators by vibrational input, in Proc. Int. Conf. Robotics and Automation,
vol. 2, May 2002, pp. 13151320.
[12] H. Yabuno, M. Miura, and N. Aoshima, Bifurcation in an inverted pendulum with tilted high-frequency excitation, J. Sound, Vibration, to be
published.
[13] J. Guckenheimer and P. Holmes, Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vector Fields. New York:
Springer-Verlag, 1983, pp. 149150.
[14] M. Golubitsuky and D. G. Schaeffer, Singularities and Groups in Bifurcation Theory. New York: Springer-Verlag, 1985, pp. 210.
[15] L. Meirovitch, Elements of Vibration Analysis. New York: McGrawHill, 1986, pp. 3031.
[16] S. A. Nayfeh and A. H. Nayfeh, The response of nonlinear systems
to modulated high-frequency input, Nonlinear Dynam., vol. 7, pp.
301315, 1995.
[17] A. H. Nayfeh, Introduction to Perturbation Techniques. New York:
Wiley, 1981, pp. 122127.
[18] J. M. T. Thompson and G. W. Hunt, Elastic Instability Phenomena. Chichester, U. K.: Wiley, 1984, pp. 4347.
[19] H. Yabuno, Y. Kurata, and N. Aoshima, Effect of Coulomb damping
on buckling of a two-rods system, Nonlinear Dynam., vol. 15, pp.
207224, 1998.
[20] T. Poston and I. Stewart, Catastrophe Theory and its Applications. London, U.K.: Pitman, 1978, pp. 7883.

AbstractA new bilateral control scheme is proposed to ensure stable


teleoperation under a wide variety of environments and operating speeds.
System stability is analyzed in terms of the time-domain definition of passivity. A previously proposed energy-based method is extended to a 2-port
network, and the issues in implementing the passivity observer and passivity controller to teleoperation systems are studied. The method is tested
with our two-degrees-of-freedom master/slave teleoperation system. Stable
teleoperation is achieved under conditions such as hard wall contact (stiff150 kN/m) and hard surface following.
ness
Index TermsPassivity controller, passivity observer, teleoperation
system, time-domain passivity.

I. INTRODUCTION
The goal of teleoperation system control is to achieve transparency
while maintaining stability (i.e., such that the system does not exhibit
vibration or divergent behavior), under any operating conditions and
for any environments. To this end, several bilateral control architectures
have thus far been developed [7], [11], [13], [22], [23], [27].
In designing the bilateral controller, a classic engineering tradeoff
between transparency and stability has been an important issue, since
transparency must often be reduced in order to guarantee stable operation in the wide range of environment impedances (for example, in
terms of stiffness of free space and hard contact). This has necessitated investigating methods to increase transparency without introducing instability. Several previous studies have sought out theoretical
design methods for control parameters, based on linear circuit theory
[1], [12] or linear robust control theory [4], [18], [26].
However, the teleoperation systems of our interest are nonlinear, and
the dynamic properties of a human operator are always involved. These
factors make it difficult to analyze teleoperation systems in terms of
known parameters and linear control theory. To cope with the nonlinearity and uncertain parameters of the teleoperation system, several researchers have used nonlinear control laws, such as adaptive control,
to design the bilateral controller [10], [17], [21], [29]. However, this
approach requires, at the very least, system dynamic equations, and the
system uncertainty should be captured with a few unknown parameters.
Generally, it is very difficult to obtain an exact dynamic model of the
teleoperation system. Furthermore, the dynamic structure of a teleoperation system is too complicated to capture with just a few parameters.
Thus, it becomes very complicated to apply this model-based approach
when the teleoperation system has high degrees of freedom (DOFs).
One promising approach is the use of the idea of passivity to guarantee stable operation without exact knowledge of model information.
Anderson and Spong [3] and Neimeyer and Slotine [20] have used
Manuscript received November 21, 2002; revised August 5, 2003. This paper
was recommended for publication by Associate Editor P. Dupont and Editor
A. De Luca upon evaluation of the reviewers comments. This work was supported in part by a grant from Ford Motor Company, and in part by the Postdoctoral Fellowship Program of the Korea Science and Engineering Foundation
(KOSEF).
J.-H. Ryu is with the Department of Electrical Engineering and Computer Science, Korea Advanced Institute of Science and Technology, Taejeon 305-701,
Korea (e-mail: jhryu@rit.kaist.ac.kr).
D.-S. Kwon is with the Department of Mechanical Engineering, Korea Advanced Institute of Science and Technology, Taejeon 305-701, Korea (e-mail:
kwonds@kaist.ac.kr).
B. Hannaford is with the Department of Electrical Engineering, University of
Washington, Seattle, WA 98195-2500 USA (e-mail: blake@u.washington.edu).
Digital Object Identifier 10.1109/TRA.2004.824689

1042-296X/04$20.00 2004 IEEE

You might also like