You are on page 1of 13

3

Robot Dynamics

Robot manipulators are articulated mechanical systems composed of links


connected by joints. The joints are mainly of two types: revolute and prismatic.
In this textbook we consider robot manipulators formed by an open kinematic
chain as illustrated in Figure 3.1.

q3
z0 zn
z2 link 2
joint n
joint 2 z3
link n
q2 qn
z1
link 1 ⎡ ⎤
x1
joint 1 ⎢x2⎥
x = ⎣ .. ⎦
q1 .
xm

y0
x0

Figure 3.1. Abstract diagram of an n-DOF robot manipulator

Consider the generic configuration of an articulated arm of n links shown


in Figure 3.1. In order to derive the mathematical model of a robot one typ-
ically starts by placing a 3-dimensional reference frame (e.g. in Cartesian
coordinates) at any location in the base of the robot. Here, axes will be la-
beled with no distinction using {x y z} or {x0 y0 z0 } or {x1 x2 x3 }. The
links are numbered consecutively from the base (link 0) up to the end-effector
(link n). The joints correspond to the contact points between the links and
60 3 Robot Dynamics

are numbered in such a way that the ith joint connects the ith to the (i − 1)th
link. Each joint is independently controlled through an actuator, which is usu-
ally placed at that joint and the movement of the joints produces the relative
movement of the links. We temporarily denote by zi , the ith joint’s axis of
motion. The generalized joint coordinate denoted by qi , corresponds to the
angular displacement around zi if the ith joint is revolute, or to the linear
displacement along zi if the ith joint is prismatic. In the typical case where
the actuators are placed at the joints among the links, the generalized joint
coordinates are named joint positions. Unless explicitly said otherwise, we
assume that this is the case.

z3
z1
z0 z2 q3

z4

q2
q1 q4
⎡ ⎤
x1
x = ⎣ x2 ⎦
y0 x3

x0

Figure 3.2. Example of a 4-DOF robot

Example 3.1. Figure 3.2 shows a 4-DOF manipulator. The placement


of the axes zi as well as the joint coordinates, are illustrated in this
figure. ♦

The joint positions corresponding to each joint of the robot, and which
are measured by sensors adequately placed at the actuators, that are usually
located at the joints themselves, are collected for analytical purposes, in the
vector of joint positions q. Consequently, for a robot with n joints, that is,
with n DOF (except for special cases, such as elastic-joints or flexible-link
robots), the vector of joint positions q has n elements:
3 Robot Dynamics 61

⎡ ⎤
q1
⎢ q2 ⎥
q=⎢ ⎥
⎣ .. ⎦
.
qn

i.e. q ∈ IRn . On the other hand it is also of great interest, specifically from a
practical viewpoint, to determine the position and orientation (posture) of the
robot’s end-effector since it is the latter that actually carries out the desired
task. Such position and orientation are expressed with respect to the reference
frame placed at the base of the robot (e.g. a Cartesian frame {x0 , y0 , z0 }) and
eventually in terms of the so-called Euler angles. Such coordinates (and angles)
are collected in the vector x of operational positions1
⎡ ⎤
x1
⎢ x2 ⎥
x=⎢ ⎥
⎣ ... ⎦
xm

where m ≤ n. In the case when the robot’s end-effector can take any position
and orientation in the Euclidean space of dimension 3 (e.g. the room where
the reader is at the moment), we have m = 6. On the other hand, if the
robot’s motion is on the plane (i.e. in dimension 2) and only the position of
the end-effector is of interest, then m = 2. If, however, the orientation on the
plane is of concern then, m = 3.
The direct kinematic model of a robot, describes the relation between the
joint position q and the position and orientation (posture) x of the robot’s
end-effector. In other words, the direct kinematic model of a robot is a function
ϕ : IRn → IRm such that
x = ϕ(q) .
Although lengthy, computation of the direct kinematic model, x = ϕ(q),
is methodical and in the case of robots of only few DOF, it involves simple
trigonometric expressions.
The inverse kinematic model consists precisely in the inverse relation of
the direct kinematic model, that is, it corresponds to the relation between the
operational posture x and the joint position q, i.e.
q = ϕ−1 (x).

In contrast to the direct kinematic model, computation of the inverse kine-


matic model q = ϕ−1 (x) may be highly complex and, as a matter of fact, may
yield multiple solutions.
The dynamic model of a robot consists of an ordinary differential equation
where the variable corresponds to the vector of positions and velocities, which
1
These are also known as positions in the operational space or workspace.
62 3 Robot Dynamics

may be in joint coordinates q, q̇ or in operational coordinates x, ẋ. In general,


these are second-order nonlinear models which can be written in the generic
form

f EL (q, q̇, q̈, τ ) = 0 , (3.1)


f C (x, ẋ, ẍ, τ ) = 0 . (3.2)

The vector τ stands for the forces and torques applied at the joints by
the actuators. The dynamic model (3.1) is the dynamic model in joint space,
while (3.2) corresponds to the dynamic model in operational space. In this
text, we focus on the dynamic model in joint space and, for simplicity, we
omit the words “in joint space”.
Kinematics as much as dynamics, is fundamental to plan and carry out
specific tasks for robot manipulators. Both concepts are dealt with in consid-
erable detail in a large number of available textbooks (see the list of references
at the end of this chapter). However, for the sake of completeness we present
in this chapter the basic issues related to robot dynamics. In the next two
chapters we present properties of the dynamic model, relevant for control. In
Chapter 5 we present in detail the model of a real 2-DOF lab prototype which
serves as a case study throughout the book.
Besides the unquestionable importance that robot dynamics has in control
design, the dynamic models may also be used to simulate numerically (with a
personal computer and specialized software), the behavior of a specific robot
before actually being constructed. This simulation stage is important, since it
allows us to improve the robot design and in particular, to adapt this design
to the optimal execution of particular types of tasks.
One of the most common procedures followed in the computation of the
dynamic model for robot manipulators, in closed form (i.e. not numerical), is
the method which relies on the so-called Lagrange’s equations of motion. The
use of Lagrange’s equations requires the notion of two important concepts
with which we expect the reader to be familiar: kinetic and potential energies.
We describe next in some detail, how to derive the dynamic model of a
robot via Lagrange’s equations.

3.1 Lagrange’s Equations of Motion

The dynamic equations of a robot manipulator in closed form may be obtained


from Newton’s equations of motion or, via Lagrange’s equations. All of these
are well documented in textbooks on analytical mechanics and are only briefly
presented here.
The disadvantage of the first method is that the complexity of the analysis
increases with the number of joints in the robot. In such cases, it is better
3.1 Lagrange’s Equations of Motion 63

to use Lagrange’s equation of motion. The latter are named after the French
mathematician Joseph Louis de La Grange (today spelled “Lagrange”) which
first reported them in 1788 in his celebrated work “Mécanique analytique”2 .
Consider the robot manipulator with n links depicted in Figure 3.1. The
total energy E of a robot manipulator of n DOF is the sum of the kinetic and
potential energy functions, K and U respectively, i.e.

E (q, q̇) = K (q, q̇) + U(q)

where q = [q1 , · · · , qn ]T .
The Lagrangian L(q, q̇) of a robot manipulator of n DOF is the difference
between its kinetic energy K and its potential energy U, that is,

L (q, q̇) = K (q, q̇) − U(q) . (3.3)

We assume here that the potential energy U is due only to conservative


forces such as gravitational energy and energy stored in compressed springs.
Then, the Lagrange equations of motion for a manipulator of n DOF, are
given by ' (
d ∂L(q, q̇) ∂L(q, q̇)
− = τ,
dt ∂ q̇ ∂q
or in the equivalent form by
' (
d ∂L(q, q̇) ∂L(q, q̇)
− = τi , i = 1, · · · , n (3.4)
dt ∂ q̇i ∂qi

where τi correspond to the external forces and torques (delivered by the ac-
tuators) at each joint as well as to other (nonconservative) forces. In the class
of nonconservative forces we include those due to friction, the resistance to
the motion of a solid in a fluid, and in general, all those that depend on time
and velocity and not only on position.
Notice that one has as many scalar dynamic equations as the manipulator
has degrees of freedom.
The use of Lagrange’s equations in the derivation of the robot dynamics
can be reduced to four main stages:

1. Computation of the kinetic energy function K(q, q̇) .


2. Computation of the potential energy function U(q) .
3. Computation of the Lagrangian (3.3) L(q, q̇) .
4. Development of Lagrange’s equations (3.4).
2
These equations are also called Euler–Lagrange equations in honor to the Swiss
scientist Leonhard Euler, contemporary of La Grange, who discovered a similar
but more general set of second-order differential equations.
64 3 Robot Dynamics

In the rest of this section we present some examples that illustrate the
process of obtaining the robot dynamics by the use of Lagrange’s equations
of motion.

z0 z0 m2

l2

m1 ϕ

l1

x0
x0 y0
y0 q1
Figure 3.3. Example of a 1-DOF mechanism

Example 3.2. Consider the mechanism shown in Figure 3.3. It consists


of a rigid link formed by two parts, of lengths l1 and l2 , whose masses
m1 and m2 are, for simplicity, considered to be concentrated at their
respective centers of mass, located at the ends. The angle ϕ is constant.
The mechanism possesses only revolute motion about the z0 axis,
the angle of which is represented by q1 . For this example, the only
degree-of-freedom is associated to the joint 1. Then, q is a scalar de-
fined as q = q1 .
We emphasize that the dynamic model of this mechanism may be
obtained using the concepts of dynamics of rigid bodies in rotation,
which are subjects of study in elementary courses of physics. However,
for the sake of illustration we employ Lagrange’s equations of motion.
The kinetic energy function K(q, q̇) of the system is given by the
product of half the moment of inertia times the angular velocity
squared, i.e.
1
K (q, q̇) = m2 l22 cos2 (ϕ) q̇ 2
2
and the corresponding potential energy

U(q) = m1 l1 g + m2 (l1 + l2 sin(ϕ)) g


3.1 Lagrange’s Equations of Motion 65

where g is the gravity acceleration. Here, we assumed that the poten-


tial energy function is zero on the plane x0 –y0 . Actually, notice that
in this example the potential energy is constant so in particular, it
does not depend on the joint position q.
The Lagrangian L(q, q̇), expressed by (3.3), is in this case
m2 2
L(q, q̇) = l cos2 (ϕ) q̇ 2 − m1 l1 g − m2 (l1 + l2 sin(ϕ)) g .
2 2
From this, one can obtain the following equations:
∂L
= m2 l22 cos2 (ϕ) q̇
∂ q̇
' (
d ∂L
= m2 l22 cos2 (ϕ) q̈
dt ∂ q̇
∂L
= 0.
∂q

Then, the corresponding Lagrange Equation (3.4) is

m2 l22 cos2 (ϕ) q̈ = τ , (3.5)

where τ is the torque applied at the joint 1. Equation (3.5) describes


the dynamic behavior of the mechanism. Notice that the dynamic
model is simply a linear second-order nonautonomous differential
equation.
Equation (3.5) may be expressed in terms of the state vector [q q̇]T
as ⎡ ⎤ ⎡ ⎤ ⎡ ⎤
q q̇ q(0)
d ⎣ ⎦ ⎢ ⎥ ⎣ ⎦ ∈ IR2 .
=⎣ τ ⎦;
dt
q̇ m2 l22 cos2 (ϕ) q̇(0)

The necessary and sufficient condition for the existence of equi-


libria is τ (t) = 0 for all t ≥ 0. In this situation, the equation has an
infinite number of equilibria which are given by [q q̇]T = [q ∗ 0]T ∈ IR2
with q ∗ ∈ IR. The interpretation of this result is the following. If at the
instant t = 0 the position q(0) has any value q ∗ ∈ IR, the velocity q̇(0)
is zero and moreover no torque is applied at the joint (i.e. τ (t) = 0 for
all t) then, we have q(t) = q ∗ and q̇(t) = 0 for all t ≥ 0. Note that the
latter is in accordance with the physical interpretation of the concept
of equilibrium. ♦

The following example illustrates the derivation of the dynamic model of


the 2-DOF robot with revolute joints shown in Figure 3.4 and that moves
about purely on the horizontal plane. Therefore, gravity has absolutely no
66 3 Robot Dynamics

influence on the robot dynamics. This should not surprise the reader; note
that the potential energy for this robot is constant since its mass does not
move in the vertical direction.
A particularity of this example, is that the actuators that deliver the
torques are not physically located at the joints themselves, but one of them
transmits the motion to the link through belts. Such types of transmission are
fairly common in industrial robots. The motivation for such configurations is
to lighten the weight (henceforth the inertia) of the arm itself by placing the
actuators as close to the base as possible.

z
τ1
x
l1 q1
I1 lc1
m1 y

τ2

q2 δ
m2
I2
lc2

Figure 3.4. Example of a 2-DOF robot

Example 3.3. Consider the robot manipulator with 2 DOF shown in


Figure 3.4. This manipulator consists of two rigid links where l1 is
the length the first link. Both joints are revolute. The robot moves
about only on the horizontal plane x–y as is shown in Figure 3.4. The
masses of the links are denoted by m1 and m2 respectively. Notice
that the center of mass of link 2 may be physically placed “out” of
the link itself! This is determined by the value of the constant angle δ.
The distances of the rotating axes to the centers of mass, are denoted
by lc1 and lc2 respectively. Finally, I1 and I2 denote the moments
of inertia of the links with respect to axis that passes through their
3.1 Lagrange’s Equations of Motion 67

respective centers of mass and that are parallel to the z axis. The joint
positions associated to the angles q1 and q2 are measured between each
respective link and an axis which is parallel to the x axis. Both angles
are taken to be positive counterclockwise. The motion is transmitted
to link 2 by a belt since the corresponding actuator is placed at the
base of the robot. The vector of joint positions q is defined as

q = [q1 q2 ]T .

The kinetic energy function K(q, q̇) for this arm may be decom-
posed into the sum of two parts: K(q, q̇) = K1 (q, q̇) + K2 (q, q̇), where
K1 (q, q̇) and K2 (q, q̇) are the kinetic energies associated with the
masses m1 and m2 respectively. In turn, the kinetic energy includes the
linear and angular motions. Thus, we have K1 (q, q̇) = 12 m1 v12 + 12 I1 q̇12 ,
where v1 is the speed3 of the center of mass of link 1. In this case,
1 1
K1 (q, q̇) = 2 2
m1 lc1 q̇1 + I1 q̇12 . (3.6)
2 2
On the other hand, K2 (q, q̇) = 21 m2 v22 + 12 I2 q̇22 , where v2 is the
speed of the center of mass of link 2. This speed squared, i.e. v22 , is
given by
v22 = ẋ22 + ẏ22
where ẋ2 and ẏ2 are the components of the velocity vector of the
center of mass of link 2. The latter are obtained by evaluating the
time derivative of the positions x2 and y2 of the center of mass of link
2, i.e.
x2 = l1 cos(q1 ) + lc2 cos(q2 − δ)
y2 = l1 sin(q1 ) + lc2 sin(q2 − δ) .
Using the trigonometric identities, cos(θ)2 + sin(θ)2 = 1 and
sin(q1 )sin(q2 − δ) + cos(q1 )cos(q2 − δ) = cos(q1 − q2 + δ), we finally get

v22 = l12 q̇12 + lc2


2 2
q̇2 + 2l1 lc2 cos(q1 − q2 + δ)q̇1 q̇2

which implies that


m2 2 2 m2 2 2
K2 (q, q̇) = l q̇ + l q̇ + m2 l1 lc2 cos(q1 − q2 + δ)q̇1 q̇2
2 1 1 2 c2 2
1
+ I2 q̇22 . (3.7)
2
Since the robot moves only on the horizontal plane, the potential
energy is constant, e.g. U(q) = 0 .
3
Some readers may be surprised that we use the word speed as opposed to velocity.
We emphasize that velocity is a vector quantity hence, it has a magnitude and
direction. The magnitude of the velocity is called speed.
68 3 Robot Dynamics

So from Equations (3.6) and (3.7), it is clear that the Lagrangian,

L(q, q̇) = K(q, q̇) − U(q) ,

takes the form

L(q, q̇) = K1 (q, q̇) + K2 (q, q̇)


1 1
= (m1 lc1 2
+ m2 l12 )q̇12 + m2 lc2
2 2
q̇2
2 2
+ m2 l1 lc2 cos(q1 − q2 + δ)q̇1 q̇2
1 1
+ I1 q̇12 + I2 q̇22 .
2 2
From this equation we obtain the following expressions:
∂L
= (m1 lc1
2
+ m2 l12 + I1 )q̇1
∂ q̇1
+ m2 l1 lc2 cos(q1 − q2 + δ)q̇2 ,
' (
d ∂L ) *
= m1 lc1
2
+ m2 l12 + I1 q̈1
dt ∂ q̇1
+ m2 l1 lc2 cos(q1 − q2 + δ)q̈2
− m2 l1 lc2 sin(q1 − q2 + δ)(q̇1 − q̇2 )q̇2 ,

∂L
= −m2 l1 lc2 sin(q1 − q2 + δ)q̇1 q̇2 ,
∂q1
∂L
= m2 lc2
2
q̇2 + m2 l1 lc2 cos(q1 − q2 + δ)q̇1 + I2 q̇2 ,
∂ q̇2
' (
d ∂L
= m2 lc2
2
q̈2 + m2 l1 lc2 cos(q1 − q2 + δ)q̈1
dt ∂ q̇2
− m2 l1 lc2 sin(q1 − q2 + δ)q̇1 (q̇1 − q̇2 ) + I2 q̈2 ,

∂L
= m2 l1 lc2 sin(q1 − q2 + δ)q̇1 q̇2 .
∂q2
Thus, the dynamic equations that model the robot manipulator
may be obtained by using Lagrange’s equations (3.4),
+ ,
τ1 = m1 lc1
2
+ m2 l12 + I1 q̈1
+ m2 l1 lc2 cos(q1 − q2 + δ)q̈2
+ m2 l1 lc2 sin(q1 − q2 + δ)q̇22

and,
3.1 Lagrange’s Equations of Motion 69

τ2 = m2 l1 lc2 cos(q1 − q2 + δ)q̈1 + [m2 lc2


2
+ I2 ]q̈2
− m2 l1 lc2 sin(q1 − q2 + δ)q̇12 .

Finally using the identities

cos(q1 − q2 + δ) = cos(δ)cos(q1 − q2 ) − sin(δ)sin(q1 − q2 )


= cos(δ)cos(q2 − q1 ) + sin(δ)sin(q2 − q1 ),
sin(q1 − q2 + δ) = cos(δ)sin(q1 − q2 ) + sin(δ)cos(q1 − q2 )
= −cos(δ)sin(q2 − q1 ) + sin(δ)cos(q2 − q1 ) ,

and denoting C21 = cos(q2 − q1 ), S21 = sin(q2 − q1 ), one obtains


+ ,
τ1 = m1 lc1
2
+ m2 l12 + I1 q̈1
+ [m2 l1 lc2 cos(δ)C21 + m2 l1 lc2 sin(δ)S21 ]q̈2
+ [−m2 l1 lc2 cos(δ)S21 + m2 l1 lc2 sin(δ)C21 ]q̇22
(3.8)

and

τ2 = [m2 l1 lc2 cos(δ)C21 + m2 l1 lc2 sin(δ)S21 ]q̈1


+ [m2 lc2
2
+ I2 ]q̈2
+ [m2 l1 lc2 cos(δ)S21 − m2 l1 lc2 sin(δ)C21 ]q̇12
(3.9)

which are of the form (3.1) with


' (
τ − RHS(3.8)
f EL (q, q̇, q̈, τ ) = 1
τ2 − RHS(3.9)

where RHS(3.8) and RHS(3.9) denote the terms on the right-hand


side of (3.8) and (3.9) respectively. ♦

In the following example we derive the dynamic model of a Cartesian robot


with 3 DOF whose main feature is that it has a linear dynamic model.

Example 3.4. Consider the 3-DOF Cartesian robot manipulator shown


in Figure 3.5. The manipulator consists of three rigid links mutually
orthogonal. The three joints of the robot are prismatic. The robot’s
displacements happen in the space x0 –y0 –z0 shown in Figure 3.5. The
vector of joint positions is q = [q1 q2 q3 ]T .
The kinetic energy function for this manipulator is given by (cf.
Figure 3.5):
70 3 Robot Dynamics

z0 q2 z0
m3 m2

q3 m1

q1
q1
y0 y0
q2
x0 x0 q3

Figure 3.5. Example of a 3-DOF Cartesian robot

1+ ,
K(q, q̇) = m1 q̇32 + [m1 + m2 ]q̇22 + [m1 + m2 + m3 ]q̇12 . (3.10)
2
On the other hand, the potential energy is given by

U(q) = [m1 + m2 + m3 ]gq1 . (3.11)

From Equations (3.10) and (3.11) we obtain the Lagrangian,

L(q, q̇) = K(q, q̇) − U(q)


1+ ,
= m1 q̇32 + [m1 + m2 ]q̇22 + [m1 + m2 + m3 ]q̇12
2
− [m1 + m2 + m3 ]gq1 .

So we have
' (
∂L d ∂L
= [m1 + m2 + m3 ]q̇1 = [m1 + m2 + m3 ]q̈1
∂ q̇1 dt ∂ q̇1
' (
∂L d ∂L
= [m1 + m2 ]q̇2 = [m1 + m2 ]q̈2
∂ q̇2 dt ∂ q̇2
' (
∂L d ∂L
= m1 q̇3 = m1 q̈3
∂ q̇3 dt ∂ q̇3

∂L ∂L ∂L
= =0 = −[m1 + m2 + m3 ]g.
∂q2 ∂q3 ∂q1
The dynamic equations that model this robot may be obtained by
applying the Lagrange’s equations (3.4) to obtain
3.2 Dynamic Model in Compact Form 71

[m1 + m2 + m3 ]q̈1 + [m1 + m2 + m3 ]g = τ1 (3.12)


[m1 + m2 ]q̈2 = τ2 (3.13)
m1 q̈3 = τ3 (3.14)

where τ1 , τ2 and τ3 are the external forces applied at each joint. Notice
that in this example Equations (3.12)–(3.14) define a set of linear
autonomous differential equations.
In terms of the state vector [q1 q2 q3 q̇1 q̇2 q̇3 ] , the equations
(3.12), (3.13) and (3.14) may be expressed as
⎡ ⎤
⎡ ⎤ q̇1
q1
⎢ ⎥
⎢ ⎥ ⎢ ⎥
⎢ ⎥ ⎢ q̇2 ⎥
⎢ q2 ⎥ ⎢ ⎥
⎢ ⎥ ⎢ ⎥
⎢ ⎥ ⎢ ⎥
⎢ ⎥ ⎢ q̇3 ⎥
⎢ q3 ⎥ ⎢ ⎥
d ⎢ ⎥ ⎢ ⎥
⎢ ⎥=⎢ 1 ⎥.
dt ⎢ ⎥ ⎢ [τ − [m + m + m ]g] ⎥
⎢ q̇1 ⎥ ⎢ m1 + m2 + m3 1 1 2 3 ⎥
⎢ ⎥ ⎢ ⎥
⎢ ⎥ ⎢ 1 ⎥
⎢ ⎥ ⎢ τ ⎥
⎢ q̇2 ⎥ ⎢ 2 ⎥
⎣ ⎦ ⎢ m 1 + m 2 ⎥
⎣ ⎦
1
q̇3 τ3
m1
The necessary and sufficient condition for the existence of equilib-
ria is τ1 = [m1 + m2 + m3 ]g, τ2 = 0 and τ3 = 0 and actually we have
an infinite number of them:
T T
[q1 q2 q3 q̇1 q̇2 q̇3 ] = [q1∗ q2∗ q3∗ 0 0 0]

with q1∗ , q2∗ , q3∗ ∈ IR. ♦

3.2 Dynamic Model in Compact Form

In the previous section we presented some examples to illustrate the applica-


tion of Lagrange’s equations to obtain the dynamic equations for robots with
particular geometries. This same methodology, however, may be employed in
general to obtain the dynamic model of any robot of n DOF.
This methodology is commonly studied in classical texts on robotics and
theoretical mechanics and therefore, we present it here only in compact form.
The interested reader is invited to see the cited texts at the end of the chapter.
Consider a robot manipulator of n DOF composed of rigid links intercon-
nected by frictionless joints. The kinetic energy function K(q, q̇) associated
with such an articulated mechanism may always be expressed as

You might also like