You are on page 1of 283

Module 1

Module 2
Module 3
Module 4
Module 5
Module 6
Module 7

Mechanical rectilinear systems


Rotational systems
Coupled mechanical systems
Nonlinear mechanical systems
Mechanical planar systems
Formulation of system equations
Formulation of transfer functions

Module 1
Mechanical rectilinear systems
Module units

1.1

Rectilinear systems and their surroundings

1.2

Rectilinear interactions

1.2.1

Rectilinear translation

1.2.2

Forces

1.2.3

1.3

1.4

1.5

Power, energy and work

Composition of rectilinear interactions

1.3.1

Composition of rectilinear motions

1.3.2

Composition of forces

Rectilinear inertia of mass

1.4.1

Pure inertor

1.4.2

Inertia energy

Prescribed rectilinear interactions

1.5.1

Prescribed velocities

1.5.2

Indicators of position and acceleration

1.5.3

Prescribed forces

1.5.4

Indicators of velocities and forces

1.6

Rectilinear damping

1.7

Rectilinear compliance

1.8

1.7.1

Rectilinear deformations

1.7.2

Spring potential energy

Summary

Module overview.
This module will introduce you to dynamic behavior of mechanical systems the bodies of which are in rectilinear translational motion,
i.e. they move in parallel straight lines. (Mechanical systems with bodies in curvilinear translation are treated in the module on planar
mechanical systems.)

Motions of various bodies in such systems are nearly always associated with coexisting forces. In many instances we prefer to think of
motion as resulting from the application of a force. In other instances we may prefer to think of a certain force as resulting from a
given motion. In either case, dynamic behavior of translational systems is determined by energy interactions that occur between
bodies and their surroundings.
When investigating dynamic behavior of translational systems it is convenient to make use of three types of idealized elements
modeling the essential dynamic phenomena which occur in various ways in bodies of these systems. It is the inertor modeling inertial
effects of a moving mass, the pure spring modeling compliance of flexible bodies, and the pure damper modeling dissipation of
mechanical energy as heat. In addition, physical elements in the form of pure sources of velocity and pure sources of force are useful
for modeling drivers imposing prescribed motions or forces on the bodies. Dynamic models of translational systems can be set up
easily from the physical elements. These models are represented graphically by multipole diagrams consisting from symbols of the
elements.
Module objectives. When you have completed this module you should be able to:
1. Understand the fundamental laws governing dynamics of mechanical translational systems.
2. Set up multipole diagrams representing dynamic models of these systems from physical elements.
3. Use the computer package DYNAST to submit multipole diagrams representing the multipole diagrams in their graphical form.
4. Use DYNAST to specify excitation and initial conditions of these models, invoke numerical time-domain analysis, and plot the
results.
Module prerequisites. High-school physics and mathematics.
1.1 Rectilinear systems and their surroundings

To define the rectilinear system of our interest, we usually begin the investigation of the system dynamic behavior with drawing a
sketch of the system geometric configuration showing dimensions and positions of the system bodies. To determine directions of the
body motions and of the related forces, we associate the system sketch with a coordinate axes.

Fig. 1.1 shows, for example, a sketch of two cars in the rectilinear motion along a straight horizontal road. A car is towing another car
using a towing rod connected between the car hooks A and B. The assumed positive direction of the car motions is in agreement with
the direction of the x axis.

Figure 1.1 Car towing.


Bodies of the cars constitute together with the body of the rod a mechanical rectilinear system. Behavior of the system is influenced by
the system surroundings formed by the wind blowing opposite to the car motions as well as by the friction between the car tyres and
the road surface.

Figure 1.2 Wheel suspension system.


Fig. 1.2 was sketched for the purpose of the dynamic analysis focused on the behavior of a car-wheel suspension. When the car travels
along a road, the wheel tyre copies potholes in the road surface. The wheel is thus driven up or down in the vertical direction along the
axis y. In this case, the rectilinear system under consideration consists of four bodies: the spring, the shock absorber, the wheel and the

quarter of the car body. The system excitation by the road surface and the gravitational attraction of the wheel and the quarter-car body
forms the system surroundings in this case.

1.2 Rectilinear interactions

1.2.1 Rectilinear translation


1.2.2 Forces
1.2.3 Power, energy and work
1.2.1 Rectilinear translation

Rectilinear motion of a point is defined as the motion of that point along a straight line. Translational motion of a body can be either
rectilinear or curvilinear as shown in Fig. 1.3. All points of a body in rectilinear translation move in parallel straight lines. This is in
contrast to curvilinear translation of bodies the points of which move on congruent curves.

Figure 1.3 Body in (a) rectilinear and (b) curvilinear translation.


If a body can be considered rigid its rectilinear motion is completely specified by the motion of any one point in the body, since all
points have the same motion. A rigid body is a body whose changes in shape are negligible compared with the overall dimensions of
the body or with the changes in position of the body as a whole. If each of the cars shown in Fig. 1.1 is considered as a rigid body,
motions of the cars can be represented, for example, by the points A and B denoting the cars rod hooks.
Absolute rectilinear motion of a given point is defined as the motion of that point relative to a fixed point in space considered as a
reference frame for the motion. Because of the difficulty of providing a real fixed point in space from which to make measurements,
the concept of absolute motion is strictly hypothetical, but nevertheless very useful.
A reference frame fixed to the earth is often a reasonable approximation. However, when dealing with the motion of long-range
missiles, for example, the earth cannot be considered to be a nonaccelerating reference and velocities are then measured with respect to
the fixed stars instead.
Fig. 1.4 shows a geometric configuration of points A and B representing the car hooks in Fig. 1.1. Motion of the points is constrained
to the rectilinear translation along the x-axis painted on the reference frame. The position xA(t) of point A is defined as the distance of
this point from the origin 0 chosen on the axis x. In the figure, xA(t) is shown negative whereas the position xB(t) of point B is positive.
Positions are measured in meters [m].

Figure 1.4 (a) Positions and (b) velocities of points in rectilinear motion.
The absolute velocity

[m/s] and the absolute acceleration

[m/s2] of point A is by definition

We shall depict the absolute velocities in drawings of geometric configurations by empty-head arrows as shown in Fig. 1.4.
Displacements, velocities and accelerations are taken as positive if they correspond to a motion in the chosen positive direction of the
related coordinate axis.
We shall often be concerned with the relative motion of two points. The relative displacement xBA, relative velocity
acceleration
of point B with respect to point A is at any instant

and relative

(1.1
)

Position, velocity or acceleration of a point with respect to another point can be measured by a device applied to the two points
across the system without braking into it. Velocity can be thus considered as an across variable, position as an integrated across
variable, and acceleration as a differentiated across variable.

1.2.2 Forces

Force is the action of one body on another. The action of a force is characterized by its magnitude, by the direction of its action, and by
its point of application. Forces are measured in physical units called newton [N]. A force tends to move a body in the direction of its
action.
Forces acting on a body are classified as either contact forces or body forces. Contact forces are generated through direct physical
contact between two bodies or between a body and a fluid. Body forces are those applied by remote action, such as gravitational,
electrostatic or electromagnetic forces. Actually both contact and body forces are distributed forces as the contact forces are in fact
distributed over the cross-section area of the contact. In the system shown in Fig. 1.1, the forces of the wind acting on the cars are
distributed over the car surfaces, while the forces in the towing rod are distributed over the rod cross-section area. If a rigid body is
constrained to rectilinear translation all the forces acting on the body can be considered as concentrated forces acting at a point.
To take a complete account of all forces acting on a body, it is essential to construct the free-body diagram by isolating the body
completely from the rest of the system and its surroundings. This diagram consists of a closed outline of the external boundary of the
body. The diagram shows all the external contact and body forces applied to the body.

Figure 1.5 (a) Free-body diagram, (b) concentrated applied force.


Fig. 1.5a shows the free-body diagram of a towed car, a part of the rectilinear system introduced in Fig. 1.1. Ft is the force exerted on
the car by the towing rod, Fw represents the distributed force of wind acting on the car surface, and Ff is the friction force acting on the
car wheels due to the roughness of the road.

As it is indicated by full-head arrows, we have chosen the following convention for the orientation of forces applied to a body. The
assumed positive direction of all applied forces is in agreement with the direction of the reference axis. Values of those applied forces
that tend to accelerate the body in this direction are positive, if an applied force tends to decelerate the body its value is negative.
When we are dealing with an external action of a force on a rigid body in rectilinear motion, the force may be applied at any point of
the body without changing its effect on the body as a whole. Thus, if the car shown in Fig. 1.5a is considered as a rigid body, all three
forces applied on it can be replaced by one concentrated force Fa acting for example at point A as shown in Fig. 1.5b. Obviously,
(1.2
)

A force F can be viewed as the time rate of the linear momentum [N.s] (called also impulse) acting in the same direction, i.e.
(1.3
)

or
(1.4
)

where 0 is the momentum at t = 0.


Forces can be measured directly by means of a calibrated spring, for example. To measure the force at a point directly, we must severe
the system at that point and insert the measuring device between the two resultant system sections. Therefore, forces are thought of as
being applied through the measuring device and hence can be considered as belonging to the family of through variables.

1.2.3 Power, energy and work

The mechanical power delivered to a moving point by the force F which acts at that point in the direction of the positive velocity v
is the product of the velocity and the component of force which lies in the direction of the velocity. If the force F and the velocity v are
collinear,
(1.5
)

Power

is defined as the rate of flow of energy or work

so that
(1.6
)

In agreement with our convention for the orientation of applied forces, we will assume that energy is consumed in the body if the signs
of the body absolute velocity and applied force are the same (both signs are positive or both are negative).
The mechanical work

done by one body on another through a point of connection is the time integral of the power:
(1.7
)

where ta and tb are the beginning and end of the time interval over which we desire to compute the work done.
The mechanical energy supplied to a body is the sum of the work done on the body by all forces acting on it

1.3 Composition of rectilinear interactions

Computed examples
Composition of motions

Computed examples
Static equilibrium of forces

1.3.1 Composition of rectilinear motions


1.3.2 Composition of forces

Composition of rectilinear motions

Purpose
Velocity, deflection, position, acceleration. Composition of velocities, deflections and accelerations. Ideal sources of velocity and ideal
indicators of velocity. Principle of motion composition.

System
Three bodies are in a simultaneous rectilinear translational motion along the x axis fixed to the railway track: a railway carriage, a
truck moving along the carriage, and a man running along the truck decking.

To see animations,
please get Adobe Flash

Sites of Interaction
A

...

railway carriage

...

truck

...

man

Task
At t = 0, the man starts running backward with respect to the x axis at the relative speed vCB = 1 m/s with respect to the truck. After
0.5 s he turns back and starts running at the same relative speed in the opposite direction. The truck relative velocity vBA with respect to
the carriage is oscillatory, while the waveform of the carriage absolute velocity vA is saw-tooth shaped as it is shown in the following
figure.

1. To verify the specification of the three velocities in the model compute their waveforms in the time interval 0
< t < tm and display them with a common scale.

2. In the same time interval compute waveforms of velocity, deflection and acceleration
o

of the railway carriage

of the truck relative to the carriage

of the man relative to the truck

3. To demonstrate the principle of motion composition, along the closed loop 0 - Ax - Bx - Cx - 0 shown below compute sums of
velocities, deflections and accelerations

(1)
(2)

Assumptions
As velocities of all bodies are given, no dynamic effects need to be considered in the model (only a kinematic model is sufficient).

Solution
Model

The x-velocities of the three bodies are represented in the model by the nodes Ax, Bx and Cx. The given motions are enforced on the
bodies by independent sources of velocities. The mans relative velocity with respect to the carriage vCB = vmeter is measured in the
model by the ideal indicator of velocity meter. The integrator and differentiatoor blocks are used to compute positions and
accelerations from velocities.
Results

Velocities vA, vBA and vCB:

Man-truck motion:

Truck-carriage motion:

Railway carriage motion:

Composition of motions along the closed loop 0 - Ax - Bx - Cx - 0:

OPEN

Explain the variations in distances between marks showing the time instances at which the corresponding
points were calculated.

Demonstrate validity of the principle of motion composition in the other loops of the simple system, e.g., in
the loop Ax - Bx - Cx, or 0 - Bx - 0.

Data

The following data were generated by DYNAST for the above given system model. You can modify the model parameters and resubmit the data to DYNAST to receive new results.

*: Composition of rectilinear motions


*SYSTEM;
tf = 1; ::[s] final time
:------------------------------------------------vA = 6*(time*(time<.5)+(1-time)*(time>.5)); ::[m/s] given carriage
vBA = 3*sin(2pi*time);
::[m/s] given truck-tovCB = 1 - 2*(time<0.5);
::[m/s] given man-to-t
source1 > E Ax = vA;
:carriage velocity drive
source2 > E Bx-Ax = vBA; :truck-to-carriage velocity driver
source3 > E Cx-Bx = vCB;
:man-to-truck velocity driver
meter > J Cx-Ax = 0;
:mann-carriage velocity
v_meter = V.meter;

Show results as

Origin

H. Mann.
[1] H. Mann: Computer-Aided Design of Dynamic Systems (in Czech). House of Engineering, Prague 1986,
p.64.

Static equilibrium of forces*

Purpose
Force, gravitational force, weight. Ideal sources of force and ideal indicators of force. Force static equilibrium.

System
A spring dynamometer that has been designed to measure vertical forces.

Sites of Interaction
A

...

dynamometer movable part

Task
The external vertical forces F1 and F2 acting on the dynamometer in the y-direction are defined as

Compute the sum of all forces Fsum acting on the dynamometer movable part in the dynamometer static steady-state.

Assumptions
Assume that the dynamometer is in a static steady-state so that the inertial and friction forces acting on the dynamometer can be
neglected.

System Parameters
m = 0.1

[kg]

mass of the dynamometer movable part

g = 9.81

[m/s]

gravitational acceleration

Solution
Model

The translatory motion of the dynamometer movable part is represented in the model by the node Ay. The dynamometer forcemeasuring effect is modeled by an ideal force meter Fmeter (i.e., an independent source of zero velocity). The gravitational force Fg =
m * g acting on the movable part is respected by the independent source of force Fg. The other two independent sources of force model
the external forces F1 and F2. The resulting sum Fsum of all forces acting at A can be computed using the explicit equation
Fsum = F1 + F2 + Fg - I.Fmeter
where I.Fmeter is the reaction force of the dynamometer. As only the static steady-state is to be investigated, the static analysis
(invoked by the DC command) can be applied to the model.
Results

Results of the static analysis are given in the file TRA1.O in the form:
Static equilibrium of forces
1

...

F1

...

F2

...

I.Fmeter

...

Fg

...

Fsum

3.500000E+000

-4.500000E+000

-9.810000E-001

-2.220446E-016

-1.981000E+000

OPEN From there we can see that the sum of all the vertical forces Fsum = 0 is zero (within the accuracy of the computer arithmetics).
Which physical law is related to this result?
Data

The following data were generated by DYNAST for the above given system model. You can modify the model parameters and resubmit the data to DYNAST to receive new results.
*: Static equilibrium of forces*
*SYSTEM;
m = 0.1;
::[kg] mass of the dynamometer movable part
g = 9.81;
::[m/s] gravitational acceleration
:-----------------------------------------------------------Fmeter > E Ay = 0;
:dynamometer model
F1 > J Ay = -3.5; :source of external force F1
F2 > J Ay = 4.5;
:source of external force F2
Fg > J Ay = m*g;
:source of gravity force
Fsum = F1 + F2 + Fg - I.Fmeter; :summation of forces
*TR; DC;
:static analysis
PRINT F1, F2,
:[N] external forces

Origin

H. Mann.
[1] H. Mann: Computer-Aided Design of Dynamic Systems (in Czech). House of Engineering, Prague 1986.

1.3.1 Composition of rectilinear motions

Let us consider again Fig. 1.4 showing a geometric configuration of points A and B constrained to the rectilinear translation along the
x-axis. According to (1.1), the relative displacement xBA, relative velocity
and relative acceleration
of point B with respect to
point A is at any instant
(1.9
)

These relations represent geometric constraints which are placed on the collinear rectilinear motions of points due to geometric
continuity of space. This is just a demonstration of the postulate of compatibility in rectilinear translation.
The first relation in (1.9) can be rewritten as
(1.1
0)

where x0B = xB. The left-hand-side of (1.10) expresses the final position of a point that has followed a closed path along the x-axis
starting from its origin 0 and continuing through the points A, B and back again to 0. It is not surprising, of course, that the sum of the
three path increments equals zero. Note, however, that differentiating (1.10) shows that also the sum of velocity increments as well the
sum acceleration increments along the close route equals zero.
1.3.2 Composition of forces

Fig. 1.6a gives an example in which point A constrained to rectilinear motion along the x-axis is acted on by forces F1(t) and F2(t)
collinear with x. The assumed positive direction of the forces is indicated by the full-head arrows. According to Newtons third law of
action and reaction, the forces acting on a point are in equilibrium, i.e., their resulting effect is zero. Thus,
(1.1
1)

which shows that the forces F1(t) and F2(t) are of the same magnitude but of the opposite direction at any instant. If F1 is considered as
an action force, F2 then forms the reaction to F1, and vice versa.

Figure 1.6 Balance of forces acting on (a) a point, (b) a body, (c) an ideal rectilinear link, (d) a configuration of ideal links.

The action of a force on a body can be separated into two effects, external and internal. Forces external to a body are of two kinds,
applied forces and reactive forces. The effects of applied forces internal to the body are the resulting internal stresses and strains
distributed throughout the material of the body. The relation between internal forces and internal strains involves the material
properties of the body and is studied in strength of materials, elasticity, and plasticity. In mechanics of rigid bodies, concern is given
only to the net external effects of forces. Fig. 1.6b shows the external force Fa applied on a car and the corresponding reaction force of
the car Fc = Fa.
The principle of transmissibility states that an external force applied at a point of a rigid body constrained to rectilinear motion may
be considered to be transmitted or to flow unchanged through the body to any other point of the body. Fig. 1.6c shows points A and B
moving along the x-axis. The points, which are interconnected by an ideal rectilinear link, are acted on in the x-direction by forces F1
and F2 external to the link. The ideal rectilinear link is by definition absolutely rigid and has no mass. The link is characterized by the
relations
(1.12)
(1.13)

The relation (1.13) is identical with (1.11). This indicates that due to the principle of transmissibility the force equilibrium holds also if
the forces F1(t) and F2(t) applied to the endpoints of an ideal rectilinear link. The second relation (1.13) follows from the assumption
that there are no link deformations due to forces acting on the link. As the link is massless it is not subjected to any inertial forces and
it neither accumulates, nor dissipates any energy.

Figure 1.7 (a) Applied and reaction forces. (b) Reaction forces flowing through the system. (c) Reaction forces in the pair of scales.
This example may be generalized for any number of points interconnected by ideal rectilinear links. Fig. 1.6c shows three points A, B
and C acted on by forces F1, F2 and F3, respectively. The points are linked to point D. All the points are in rectilinear translation along
the x-axis. The three forces are in equilibrium and the point velocities are identical, i.e.,
(1.14)
(1.15)

Fig. 1.7a shows two cars interconnected by a towing rod. The cars are considered as rigid bodies and the rod as an ideal rectilinear
link. The assume positive orientation of the car reaction forces Fc1 and Fc2 shown within the car outlines opposes the x-direction. The
car reaction forces can be imagined as flowing through the two-car assembly, and they are in balance, i.e.
(1.1

6)

To prove (1.16), let us consider Fig. 1.7b showing free-body diagrams of the cars and towing rod. The forces denoted outside the cars
and the rod are applied forces the assumed positive direction of which is the x-direction. Let us assume that

where Fr1 and Fr2 are forces applied on the cars by the towing rod, and Fo1 and Fo2 are the other external forces applied on the cars.
When submitting these relations into (1.16) after taking into consideration that Fc1 = Fa1 and Fc2 = Fa2 we obtain

The forces Fr1 and Fr2 acting on the rod are in balance, and the other external forces Fo1 and Fo2 applied to the system of rigid bodies
must be in balance too.
The reaction forces in the vertical links of the pair of scales shown in Fig. 1.7c must be balanced, i.e.

This conclusion can be viewed as a demonstration of the postulate of continuity in rectilinear translation.

1.4 Rectilinear inertia of mass

1.4.1 Pure inertor


1.4.2 Inertia energy

1.4.1 Pure inertor

Components of mechanical systems having mass resist changes of their velocity due to inertia forces acting on the mass. Let us
consider a body of a constant mass m [kg] in a rectilinear motion of absolute acceleration a [m/s2]. According to Newtons second law,
the inertia force F [N] acting on the mass in the opposite direction to the mass acceleration is
(1.1
7)

where v [m/s] is the absolute velocity of the mass with respect to a nonaccelerating reference frame.
Using the definition of force (1.3), Newtons second law (1.17) can be expressed also as
(1.1
8)

where is the linear momentum acting on the mass. By definition, = 0 when v = 0. Thus
(1.1
9)

The inertia effects encountered in rectilinear translation of bodies can be modelled by the idealized physical element called pure
inertor. The inertor ignores, however, damping, spring or any other energy effects in bodies. Our symbol for the inertor is shown in
Fig. 1.8a. The symbol consists of a square representing the mass and of a short line segment called a pin. The pin denotes the pole of
the inertor which represents the inertor energy interactions with the rest of the system.

Figure 1.8 (a) Symbol of inertor. (b) Momentum vs. velocity characteristic of ideal inertor. (c) Sliding body, (d) its model.
The pure inertor is associated with two variables: the inertor force F representing the inertia force of the modelled mass, and the
inertor velocity v representing the absolute velocity of the mass. The assumed positive orientation of F and v with respect to the
inertor symbol is indicated in Fig. 1.8a by the full-head and empty-head arrow, respectively.
Dynamic behavior of the pure inertor is completely described by the constitutive relation
(1.2
0)

where f(.) is a function. A pure inertor characterized by the linear constitutive relation (1.18) or (1.17), where the mass m is constant, is
called the ideal inertor. The relationship (1.18) is plotted in Fig. 1.8b.

From the theory of relativity we know that mass of a given body increases with the body velocity approaching the velocity of light. As
the present considerations are limited to velocity levels small compared with the velocity of light, Newtons second law can be applied.
No other nonlinear inertia effects are known. In some engineering problems, however, the mass of a body changes with time.
Fig. 1.8c shows the free-body diagram of a rigid body sliding with a negligible friction along a horizontal surface in the x-direction.
Let us assume that point A was chosen to represent the body rectilinear motion as well as the site of application of all the external
forces Fext acting on the body. Such a point can be considered as the body mass point, i.e. a point of the body associated with the
lumped body mass constrained to the rectilinear motion.
Fig. 1.8d gives the dynamic model of the sliding body in the form of a multipole diagram. Point A of the real body moving in the xdirection as shown in Fig. 1.8c is represented in the diagram by the node Ax. Diagram nodes represent sites of energy interaction
between bodies.
This is in agreement with our convention about the orientation arrows for reactions to external forces applied to bodies. The emptyhead arrow, which denotes the assumed positive orientation of the inertor velocity v, is directed towards the inertor pole away from the
reference frame.
As we assume that the body is not subjected to any other effects except its own inertia, there is only one physical element in the
diagram the inertor which is attached to node Ax. If no external force Fext is acting on the body in this example, the inertia force Fi
is zero and the body velocity is also zero or constant.
The inertor is in fact a two-pole physical element. But its second pole, with respect to which the inertor velocity should be measured,
must be always attached to the reference frame. For this reason, the second pole need not be shown in the inertor symbol just to
simplify the diagrams.

1.4.2 Inertia energy

From (1.17) it is obvious that the power entering an ideal inertor at a given instant is
(1.2
1)

If
, energy is being stored in the inertor, and if
, the stored energy is being retrieved from it. Which of these two states is
occurring depends both on the relative velocity and acceleration at the moment.
Integration of (1.21) gives the expression for the energy accumulated in the ideal inertor

which is called rectilinear kinetic energy. The energy is represented by the area in Fig. 1.8b between the axis and the -v
characteristic.
We can see that this energy depends directly on the velocity acquired by the inertor, but not on its force. Hence the inertor stores
energy by virtue of its velocity. The inertor energy is always positive and does not depend on the sign of either its velocity or
acceleration.
1.5 Prescribed rectilinear interactions

1.5.1 Prescribed velocities

1.5.2 Indicators of position and acceleration


1.5.3 Prescribed forces
1.5.4 Indicators of velocities and forces

1.5.1 Prescribed velocities

If a body is driven by a prescribed velocity v(t) [m/s], the actuator driving the body can be modelled by the pure source of velocity.
This idealized physical element is characterized by the constitutive relation
(1.2
2)

where f(t) is a known function of time.

Figure 1.9 (a) Symbol of pure source of velocity. (b) Actuator of prescribed relative velocity and (c) its multipole model.
Fig. 1.9a gives our graphical symbol for the pure source of velocity. The two pins sticking out of the circle denote the source poles
representing the energy interaction of the source with the rest of the system. Besides the prescribed velocity v(t), the source is
associated with its force F(t). This force counterbalances the forces imposed on the source by the system in which the source operates.

The source polarity denotes the + sign by one of its poles. The assumed positive orientation of v(t) and F(t) in relation to the symbol
polarity is indicated in Fig. 1.9a by the empty-head and full-head arrow, respectively.
Fig. 1.9b shows a real rectilinear actuator of negligible mass acting between the interaction sites A and B so that
where
va(t) is the actuator velocity. Let us assume that va(t) is sufficiently independent from the external forces Fext and Fext exerted on the
actuator by the system. Then the actuator can be modelled by a pure source of velocity as shown in Fig. 1.9c. The rectilinear motion of
the interaction sites A and B in the x-direction is represented in the model by nodes Ax and Bx.
Fa and Fa are the internal actuator forces counterbalancing the external forces. Let us assume that the source symbol modeling the
actuator is placed in the diagram in such a way that the orientation of its velocity v is in agreement with the velocity va of the actuator.
Then the source force F represents the internal actuator force Fa acting at the site of interaction represented by the node to which the +
pole of the source is attached.
The power consumed by the source of velocity is
(1.2
3)

If the velocity v and force F of the source are either both positive or both negative the modelled actuator consumes energy supplied by
the remaining system. In the former case, the system forces the actuator to extend its length whereas in the latter case the actuator
length is shortened by the system. If, however, the signs of the source velocity v and force F are different the modelled actuator
delivers energy into the system. If both signs are positive the actuator pushes the system sites apart, if both signs are negative the sites
pulled closer to each other.
Example:

Figure 1.10 (a) Car towing, (b) its model. (c) Towing velocity profile, (d) corresponding force exerted by towing car, (e) energy
accumulated in towed car, (f) position of both cars.
The most trivial dynamic model of the two-car system shown in Fig. 1.10a is given in Fig. 1.10b. As it is assumed that all the system
bodies are rigid and the towing rod is massless a single point the point A is sufficient for representing the system rectilinear motion
along x. Only inertia is considered in the towed car dynamics and the towing car is modelled by a source of velocity.
Let us consider the case when the velocity profile of the towing car velocity vc2 is triangular (Fig. 1.10c). The source force Fc2 response
(Fig. 1.10d) is then negative constant when the velocity increases, and positive constant when the velocity decreases. Therefore, the
towing car supplies energy into the system in the former case whereas it consumes energy in the latter case. In both cases, the energy
accumulated in the towed car is positive (Fig. 1.10e).
The forces Fc1 and Fc2 are in balance, i.e.
(1.2

4)

In conformity with our conventions, the arrows indicating orientation of the forces Fc1 and Fc2 oppose the x-direction in the system
geometric representation, while they are directed away from the non-reference node Ax in the model of the system.

1.5.2 Indicators of position and acceleration

Unlike velocities or forces, positions and accelerations are not primary variables representing rectilinear motion in multipole diagrams.
Nevertheless, they can be indicated very easily using blocks: integrators for positions and differentiators for accelerations.

Figure 1.11 Indication of (a) position and (b) acceleration.


To demonstrate this, Fig. 1.11a shows indication of the car position xA in the car-towing model given in Fig. 1.10b. The input of an
integrator is attached to the node Ax. Thus the velocity
associated with this node is integrated to obtain at the integrator output the
position xA(t) of the point A representing the car motion:

The initial condition of the integration xA(t0) corresponds to the initial position of point A at t = t0. The car position is for the triangular
velocity profile shown in Fig. 1.10f.

In a similar way, a differentiating block can be used to indicate the car acceleration. This illustrates Fig. 1.11b in which the
differentiator is attached to the node Ax. The indicated acceleration

is denoted at the differentiating block output as dAx.


Note that attaching a block input to a node does not disturb the force balance at the node as no force enters any block.

1.5.3 Prescribed forces

To model actuators of prescribed forces we need to introduce a pure source of force. In general, such a source exhibits a force F(t)
characterized by the constitutive relation
(1.2
5)

where f(t) is a known function of time.


Our graphical symbol for the source of force is given in Fig. 1.12a. The full-head and empty-head arrows indicate the assumed positive
orientation of the force F(t) and velocity v(t) associated with the source. The variable orientation is related to polarity of the source
symbol denoted by the + sign. The relative velocity v(t) of the source represents the velocity imposed on the force source by the rest of
the system in which the source operates.

Figure 1.12 (a) Symbol of pure source of force. (b) Actuator of prescribed force and (c) its multipole model.
Fig. 1.12b shows a real rectilinear actuator of a prescribed force Fa acting along the x-axis between the interaction sites A and B of a
system. The actuator force Fa is counterbalancing the external force Fext to the actuator. The effect of the real actuator is modelled in
Fig. 1.12c using the pure source of force. The motion of the interaction sites A and B in the x-direction is represented in the model by
the nodes Ax and Bx.
Polarity of the full-head arrow for the force Fa indicates that if Fa > 0 the actuator pulls the sites A and B closer. If Fa < 0 the actuator
pushes the sites A and B apart.
The source symbol should be oriented between the nodes in agreement with the orientation of the x axis associated with the real
actuator. Therefore, the + pin of the symbol should be attached to the node Bx representing the actuator site of interaction B that is
assumed to move faster then its site A. If the source is oriented properly its force F represents the prescribed force Fp acting at the site
B related to the + pole of the source.
If the source symbol modeling the actuator is placed in the diagram in such a way that the orientation of its velocity v is in agreement
with the velocity va of the actuator, the source force F represents the actuator force Fa at the actuator end which is related to the + pole
of the source.
If the source velocity v and force F are either both positive or both negative the modelled actuator consumes energy supplied by the
remaining system. In the former case, the system forces the actuator to extend its length whereas in the latter case the actuator length is
shortened by the system. If, however, the signs of the source velocity v and force F are different the modelled actuator delivers energy

into the system. If both signs are positive the actuator pushes the system sites apart, if both signs are negative the sites pulled closer to
each other.
Examples:

Figure 1.13 Car towing.


The free vertical fall of a ball shown in Fig. 1.14a is modelled in Fig. 1.14b using a source of force Fg to model the gravitational force
Fg = 9.81m acting on the body. The ball air resistance is neglected in the model.

Figure 1.14 (a) Free fall of ball, (b) its model.

1.5.4 Indicators of velocities and forces

Pure sources of zero force can be used as physical elements modeling ideal indicators of velocity, i.e. instruments measuring relative
velocity of distinct sites of interaction. Our graphical symbol for such an indicator is given in Fig. 1.15a. Direction of the arrow
showing the assumed positive polarity of the indicated velocity is fixed to the polarity of the symbol.

Fig. 1.15b shows this symbol utilized to indicate the relative velocity of a force actuator modelled by the pure source of force placed
between nodes Ax and Bx.

Figure 1.15 (a) Symbol of ideal indicator of velocity. (b) Indication of a source-of-force velocity.
Pure sources of zero velocity can be utilized as physical elements modeling ideal measuring instruments or indicators of force. Our
graphical symbol for such an indicator is given in Fig. 1.16a. Direction of the arrow showing the assumed positive polarity of the
indicated force is fixed to the polarity of the symbol. Fig. 1.16b shows this symbol utilized to indicate the force of a source of velocity
representing the force exerted on a velocity actuator by the system in which it operates.
Note the difference between measuring velocity and force. The direct measurement of forces requires disconnecting the system and
connecting it back by a force indicator. Velocities (as well as positions or accelerations) are measured by indicators applied to systems
without any need to disconnect them. This is the reason why velocities represent rectilinear across variables while forces are
rectilinear through variables.

Figure 1.16 (a) Symbol of ideal indicator of force. (b) Indication of a source-of-velocity force.

Body driven by given force*

Purpose
Modeling of prescribed rectilinear motion of a single body. Motion of the mass point, velocity, position, displacement, acceleration and
inertia.

System
Body constraied a translational rectilinear motion, is driven by a given time-variable force.

To see animations,
please get Adobe Flash

Sites of Interaction
A

...

driven body

Task
Set up the system model and compute the following variables:

force of the driver

velocity exerted on the body by the driver

velocity and the body position

position of the body position

acceleration the body force exerted on the body by the drivery

inertia on the body

The profile of the driver velocity IDRIVER is shown below.

Assumptions
The driver can be considered as ideal in the sense that it can deliver any force necessary for imposing the prescribed force on the body.

System Parameters
m=2

[kg]

body mass

Fm = 10

[N]

force factor

tf = 1

[s]

final time

Solution
Model

Node Ax is associated with the x-velocity of the point A representing rectilinear motion of the body. The body inertia is modelled by
the inertor MASS. The body is driven by the ideal source of force DRIVER. The integrator INT integrates the body velocity to obtain
the body position.

Results

OPEN
1. Explain the relationships between

the velocities of the driver and the body

the body velocity and the body position

the body velocity and the body acceleration

the body acceleration and the body inertia

the body inertia and the force exerted by the driver

2. Repeat the above tasks for a larger value of vm.


3. Repeat the above tasks for a larger value of m.

When explaining the replationships consider each of the three segments of the time interval individually.
Data

The following data were generated by DYNAST for the above given system model. You can modify the model parameters and resubmit the data to DYNAST to receive new results.

*: Body driven by given force*


*SYSTEM;
m = 2; ::[kg] body mass
Fm = 10; ::[N] force factor
tf = 1; ::[s] final time
:------------------------------------------:specification of given velocity by table:
F /tab/ 0,Fm, .2*tf,Fm, .2*tf,0, .4*tf,0, .4*tf,-Fm, tf,-Fm;
DRIVER > J Ax = -F(time);
MASS > C Ax = m;
:model of body inertia
BI > @Int Ax,xA;
:velocity integration
*TR; TR 0 tf;
:transient analysis

Show results as

Submodels

Integrator block

Origin

H. Mann
[1] H. Mann: Computer-Aided Design of Dynamic Systems (in Czech). House of Engineering, Prague 1986.

ntegrator block

Description
Outputs the integral of scaled input signal

where t0 is the starting time of transient analysis.

Symbol

Interface
in

input

out

output

External Parameters
c=1

Data
:: Integrator block
Int
in, ::input
out/ ::output
c = 1; ::coefficient
BI out = c*in;

coefficient

EO@;

Body driven by given velocity*

Purpose
Modeling of prescribed rectilinear motion of a single body. Motion of the mass point, velocity, position, displacement, acceleration and
inertia.

System
Body, constraied to a translational rectilinear motion, is driven by a given time-variable velocity.

To see animations,
please get Adobe Flash

Sites of Interaction
A

...

driven body

Task
Set up the system model and compute the following variables:

velocity of the driver

velocity and the body position

position of the body position

acceleration the body

inertia on the body

force exerted on the body by the driver

The profile of the driver velocity vDRIVER is shown below.

Assumptions
The driver can be considered as ideal in the sense that it can deliver any force necessary for imposing the prescribed force on the body.

System Parameters
m=2

[kg]

body mass

vm = 10

[m/s]

velocity factor

tf = 1

[s]

final time

Solution
Model

Node Ax is associated with the x-velocity of the point A representing rectilinear motion of the body. The body inertia is modelled by
the inertor MASS. The body is driven by the ideal source of velocity DRIVER. The integrator INT integrates the body velocity to
obtain the body position.

Results

OPEN
1. Explain the relationships between

the velocities of the driver and the body

the body velocity and the body position

the body velocity and the body acceleration

the body acceleration and the body inertia

the body inertia and the force exerted by the driver

2. Repeat the above tasks for a larger value of vm.


3. Repeat the above tasks for a larger value of m.

When explaining the replationships consider each of the three segments of the time interval individually.
Data

The following data were generated by DYNAST for the above given system model. You can modify the model parameters and resubmit the data to DYNAST to receive new results.

*: Body driven by given velocity*


*SYSTEM;
m = 2; ::[kg] body mass
vm = 10; ::[m/s] velocity factor
tf = 1; ::[s] final time
:------------------------------------------:specification of given velocity by table:
v /tab/ 0,0, .2*tf, vm, .4*tf, vm, tf,-2*vm;
DRIVER > E Ax = v(time);
MASS > C Ax = m;
:model of body inertia
BI > @Int Ax,xA;
:velocity integration
*TR; TR 0 tf;
:transient analysis

Show results as

Submodels

Integrator block

Origin

H. Mann
[1] H. Mann: Computer-Aided Design of Dynamic Systems (in Czech). House of Engineering, Prague 1986, p.
71.
Vertical throw of a ball*

Purpose
Gravitational force. Kinetic and potential energy of body. Energy equilibrium. Length of body path. Initial conditions.

System
A ball is thrown upwards along the vertical axis y.

To see animations,
please get Adobe Flash

Sites of Interaction
A

...

body of the ball

Task
The ball is acted on by the earth gravitational force. The initial altitude of the ball is yA(0) = 1m, its initial velocity is v(0) = 50m/s.
Compute the time variations of the ball altitude and velocity, length of the path travelled by the ball and ball acceleration for
Compute also the potential, kinetic and total energy of the ball in this interval.

Assumptions
The air resistance can be neglected.

System Parameters
m=1

[kg]

ball mass

g = 9.81

[m/s2]

gravitational acceleration

yA0 = 1

[m]

ball initial altitude

Ay0 = 50

[m/s]

ball initial velocity

tf = 10

[s]

final time

Solution
Model

s.

Node Ay represents the ball y-motion, ball inertia is modeled by the inertor MASS, and the gravitational force acting on the ball is
represented by the independent source of force Fg. The ball altitude is evaluated by integrating the ball y-velocity. Another integrator
evaluates the length of the path travelled by the ball.
Results

Since energy disipation is neglected, the ball acceleration and its total energy remains constant during the experiment, its potential and
kinetic energies have extremes at the highest point of the ball trajectory whereas the total path travelled by the ball increases
monotonously.

Data

The following data were generated by DYNAST for the above given system model. You can modify the model parameters and resubmit the data to DYNAST to receive new results.
*: Vertical throw of a ball*
*SYSTEM;
m = 1;
::[kg] ball mass
g = 9.81;
::[m/s^2] gravitational acceleration
yA0 = 1;
::[m] ball initial altitude
Ay0 = 50;
::[m/s] ball initial velocity
tf = 10;
::[s] final time
:--------------------------------------------------------mass > C Ay = m;
:model of ball inertia
Fg > J Ay = m*g;
:source og gravity force
BI1 > @Int Ay,yA;
:integration of ball altitude
BI2 > @Int 1,path;
:integration of ball path length

Show results as

Submodels

Integrator block

Origin

H. Mann
[1] H. Mann: Computer-Aided Design of Dynamic Systems (in Czech). House of Engineering, Prague 1986.

Body driven by a frequency-variable force

Purpose
Modeling of frequency variable excitation.

System
A rotating machinery operating at variable speed is subjected to a vertical exciting force which changes frequency with time.

To see animations,
please get Adobe Flash

Sites of Interaction
A

...

machinery

Task
Consider the machinery subjected to a force F(t) of constant amplitude Fa at frequences increasing linearly from 0 to 5 Hz in 0.6 s as
shown in Fig. (b). Compute the force F(t) and the machinery y-acceleration for
s.

Assumptions
Gravitation is irrelevant for this taste.

System Parameters
Fa = 890

[N]

force amplitude

m = 18

[kg]

body mass

k = 15.8k

[N/m]

spring stiffness

d = 52

[N.s/m]

damping factor

tf = 2

[s]

final time

Solution
Model

The y-motion of the machinery body is represented by the node Ay. The body inertia and its fixing to the frame is modeled by the
elements MASS, FRIC and SPRING, respectively. The exciting force Fexc(t) = Fa sin (t) is introduced into the model by the source
Fexc. The body y-acceleration is computed as the variable VD.Ay. Using blocks, the angle (t) is computed as

Results

Data

The following data were generated by DYNAST for the above given system model. You can modify the model parameters and resubmit the data to DYNAST to receive new results.

*: Body driven by a frequency-variable force


*SYSTEM;
Fa = 890; ::[N] force amplitude
m = 18; ::[kg] body mass
k = 15.8k; ::[N/m] spring stiffness
d = 52; ::[N.s/m] damping factor
tf = 2;
::[s] final time
:--------------------------------------------BS f = 5*time/0.6; :variable frequency
BI > @Int f,phi / 2pi; :variable angle
Fexc > J Ay = Fa*sin(phi);
mass > C Ay = m;

Show results as

Submodels

Integrator block

Origin

H. Mann by [1]
[1] Levy S., Wilkinson J.P.D.: The Component Element Method in Dynamics. McGraw-Hill 1976, p. 7071.
[2] H. Mann: Computer-Aided Design of Dynamic Systems (in Czech). House of Engineering, Prague 1986.

1.6 Rectilinear damping

Almost every mechanical motion is associated with some inherent degree of friction which resists the motion. Sometimes friction is
unwanted but must be tolerated and accounted for, as, for example, in bearings or in the aerodynamic drag on a moving body. In other
cases friction is desired and is designed into equipment, as in brakes, clutches, belt drives, and wedges. Wheeled vehicles depend on
friction for both starting and stopping, and ordinary walking depends on friction between the shoe and the ground. Friction is utilized
in vibration dampers and shock absorbers such as dashpots used in car suspension systems to damp car oscillations caused by riding
over bumps in the road.

Figure 1.17 (a) Symbol of pure damper, (b) F v characteristic of the ideal damper.
Fig. 1.17a gives our graphical symbol for pure rectilinear dampers. A pure rectilinear damper exerts a resistive force F that is a
function of the relative velocity v of its motion, but exhibits no mass or spring effects. Thus,
(1.2
6)

where f(.) is a function such that F = 0 when v = 0. In the case of an ideal rectilinear damper this constitutive relation is linear, i.e.,
(1.2
7)

where b [N.s/m] is the damping constant of the ideal damper. The ideal damper characteristic is shown in Fig. 1.17b.

Characteristics close to those of an ideal damper demonstrate, for example, the sliding surfaces shown in Fig. 1.18a if they are
separated by a thin layer of a fluid lubricant, or the dashpot in Fig. 1.18c if it does not move too fast. In such cases the friction forces
result from the viscous drag of a fluid under laminar flow conditions.
Fig. 1.18a shows sliding surfaces in contact pushed towards each other by some forces Fn normal to the motion. The frictional forces
Fr resisting a relative motion of the surfaces must be compensated for by external forces Fext.

Figure 1.18 (a) Sliding surfaces in contact, (b) corresponding model. (c) Oil-filled shock absorber.
In Fig. 1.18b, the friction effect between sliding surfaces is modelled by an ideal damper. Nodes Ax and Bx in the model represent xmotions of points A and B of the real surfaces in the x-direction. If the surfaces are well lubricated by a fluid, the damping constant
(1.2
8)

where A is the common area and d is the thickness of the lubrication film between the surfaces, is the viscosity coefficient of the
lubricant.

Note the way in which the polarities of the damper relative velocity v and resistive force F are associated with the damper symbol
shown in Fig. 1.17a. As the symbol is asymmetric, no + sign is necessary to denote the symbol polarity (as it was the case of pure
sources). If the damper symbol is placed in a multipole diagram in such a way that its velocity v represents polarity of the relative
velocity
, then the damper force F represents the resistive force Fr at the faster site of interaction. The damper force F thus
corresponds in Fig. 1.18b to the resistive force counterbalancing the net external force Fext acting on point B in the real bodies.
Since no mass is associated with the surfaces, the external forces Fext applied to the interaction sites A and B are balanced, and the
same must be true about the reaction forces Fr at these sites. Thus an external force applied to point B is assumed to flow through the
damper to point A without any change.
The shock absorber shown in Fig. 1.18c consists of a piston inside an oil-filled cylinder in which fluid is forced through a narrow
orifice in the piston by relative motion, thus producing resisting force. In this case, the damping constant can be approximated as
(1.2
9)

where D is the cylinder diameter, d is the diameter of the orifice, and is the length of the orifice.
The mechanical power dissipated in the pure damper as heat is

Note that it is always positive regardless of the direction of the damper relative motion.

Viscous friction experiment*

Purpose
Modeling of damping and energy dissipation.

System
A block is sliding along a horizontal plane with a well lubricated surface.

To see animations,
please get Adobe Flash

Sites of Interaction
A

...

body of the block

System excitation
Initial x-velocity of the block is

m/s.

Task
For

s, analyze the time variation of

block displacement from its starting position

friction force acting on the block

dissipated power and energy during the block motion

Assumptions
Friction between the block and the surface is viscous.

System Parameters
m=1

[kg]

block mass

d=5

[N.s/m]

viscous damping coefficient

Solution
Model

Friction between the body, the inertia of which is modeled by the inertor MASS, is represented by the damper FRIC. The body position
yA is computed by integrating the body velocity Ay.

Results

Data

The following data were generated by DYNAST for the above given system model. You can modify the model parameters and resubmit the data to DYNAST to receive new results.

*: Viscous friction experiment*


*SYSTEM;
m = 1; ::[kg] block mass
d = 5; ::[N.s/m] viscous damping coefficient
:-------------------------------------------MASS > C Ax = m;
:block inertia model
FRIC > G Ax = d;
:block friction model
Fd = I.FRIC;
:friction force
BI1 > @Int Ax,xA;
P = V.fric * I.fric;
BI2 W = P;
*TR; TR 0 .5;
:analysis time interval

Show results as

Submodels

Integrator block

Origin

H. Mann
[1] H. Mann: Computer-Aided Design of Dynamic Systems (in Czech). House of Engineering, Prague 1986

Two-mass model of car

System
A quarter car is moving along a surface.

To see animations,
please get Adobe Flash

Sites of Interaction
A

...

quarter of car body

...

wheel axis

...

wheel contact with road

System excitation
The surface of the road is approximated by the function yA = 0.1 sin (2.5x) m while the assumed car x-velocity is

m/s.

Task
Compute acceleration of the car body, of the undercarriage, and of the wheel.

System Parameters
mA = 300

[kg]

quarter of body mass

mB = 100

[kg]

mass of wheel

kAB = 10k

[N/m]

stiffness of suspension spring

kBC = 60k

[N/m]

stiffness of wheel tyre

dAB = 150

[N.s/m]

damping of suspension damper

dBC = 300

[N.s/m]

damping of wheel tyre

tf = 2

[s]

final time

Solution
Model

The vertical displacement imposed on the car wheel due to curvy surface is yC = 0.1 cos 5t m. The vertical velocity given to the wheel
modeled by the source of velocity is vCy = dyC/dt. Such differentiation can be performed by DYNAST in a semisymbolic way (denoted
by the symbol %).

Results

Data

The following data were generated by DYNAST for the above given system model. You can modify the model parameters and resubmit the data to DYNAST to receive new results.

*: Tw o-mass model of car


*SYSTEM;
mA = 300;
::[kg] quarter of body mass
mB = 100;
::[kg] mass of w heel
kAB = 10k;
::[N/m] stiffness of suspension s
kBC = 60k;
::[N/m] stiffness of w heel tyre
dAB = 150;
::[N.s/m] damping of suspension
dBC = 300;
::[N.s/m] damping of w heel tyre
tf = 2;
::[s] final time
:---------------------------------------------------------bodyA > C Ay = mA;
springAB > L Ay-By = 1/kAB;

Show results as

Submodels

Integrator block

Origin

H. Mann
1.7 Rectilinear compliance

1.7.1 Rectilinear deformations


1.7.2 Spring potential energy

1.7.1 Rectilinear deformations

Materials undergo a deformation when they are compressed or stretched. If a mechanical component experiences a steady deformation
when acted on by steady forces, it is exhibiting compliance. The component compliance can be modelled by pure springs. As these
physical elements ignore any inertial and dissipative effects, their deformation is a function only of the applied forces. Fig. 1.19a gives
our symbol of pure springs. The symbol resembles a wire coil spring shown in Fig. 1.19b, the most familiar mechanical device design
to exhibit compliance.

Figure 1.19 (a) Symbol of pure spring. (b) Coil spring stretched and compressed. (c) Ideal-spring deformation characteristic.
The spring deformations due to compliance are often approximated to be proportional to the applied force. A linear pure spring, which
is called an ideal spring, has a proportional relation between deformation and force,
(1.3
0)

where
[m] is the spring deformation, F [N] is the spring restoring force, and the constant parameter L [m.N1] is the compliance
factor of the spring. The reciprocal value of the compliance factor, K = 1/L [N.m1], is the spring stiffness termed also the spring
constant.

For example, the compliance factor of a coil spring can be derived using the methods of applied mechanics to be approximately

where D [m] is the spring pitch diameter (mean coil diameter), d [m] is the diameter of the wire, n [] is the number of coil turns, and
G [N.m2] is the shear modulus of the coil material.
The compliance factor of a cylindrical bar made of homogenous ideally elastic material is according to Hooks law
(1.3
1)

where [m] is the force-free length of the rod, A [m2] is the rod cross-sectional area, and E [N.m2] is the Youngs modulus of elasticity
of the rod material.
Fig. 1.19b shows a stretched, force-free and compressed spring. The spring deformation
and is the free length of the spring. If
the spring is stretched, if
the
characteristic of the ideal spring.

where is the spring actual length


the spring is compressed. Fig. 1.19c shows

By differentiating (1.30), we can arrive at the constitutive relation of an ideal rectilinear springs
(1.3
2)

where

is the spring relative velocity. Alternatively, the relation (1.32) can be expressed in the integral form

(1.3
3)

Fig. 1.20a shows a coil spring placed between the interaction sites A and B in the x-motion. If the spring mass is negligible in
comparison with masses of the interacting components it can assumed that the external forces Fext applied to the spring end-points are
balanced. This implies that also the restoring forces Fr, opposing the external forces, are balanced at the spring end-points.

Figure 1.20 (a) Coil spring, (b) force-controlled and (c) elongation-controlled model. (d) Model respecting spring mass and damping.
(e) Distributed-parameter model.
Note that one of the pins of the pure-spring symbol given in Fig. 1.19b is denoted by the + sign. This allows for an unambiguous
association of the polarities of the spring velocity v and the spring force F with the polarity of the symbol. In Fig. 1.20b, the purespring symbol is oriented in conformity with the chosen orientation of the modelled real spring shown in Fig. 1.20a. Then the purespring force F represents the real-spring restoring force Fr at the assumed faster end of the real spring.
To express the total length of the modelled spring, the pure spring is combined with an integrator in Fig. 1.20b. For a chosen preloading force F(t0) of the spring, the initial relative distance between the spring end-points at t = t0 should be specified as

(1.3
4)

where

is the initial length of the spring.

Fig. 1.20c gives an alternative elongation-controlled model to the ideal rectilinear spring shown in Fig. 1.20b. The ideal spring is
replaced there by a source of the restoring force
(1.3
5)

The source is controlled by the relative distance of the spring end points xBA. The initial value of this distance for a chosen initial length
of the spring
is
(1.3
6)

Figure 1.21 (a) Suspended ball, (b) its dynamic model, (c) ball position.
Examples:

Fig. 1.21c shows a dynamic model of the ball suspended on a spring given in Fig. 1.21a. Node Ay represents the y-motion of point A of
the ball. The spring symbol is placed between the node Ay and the reference node in such a way that the polarity of its velocity v is in
agreement with the polarity of the relative velocity of the real spring along the y axis. As a result, the pure spring force F represents,
both by its magnitude and polarity, the actual restoring force counterbalancing the gravitational force Fg acting on point A in the real
bodies. The effect of the gravitational force Fg = 9.81m is modelled using a source of force.

Figure 1.22 (a) Car following bumpy road, (b) car-wheel suspension, (c) quarter-car model.
Fig. 1.22a shows a car following a bumpy road while Fig. 1.22b shows just a detail of the car, the suspension of one of its wheels. In
Fig. 1.22c point Ry represents the y-motion of the contact point between the car tyre and the road surface R when the car moves in the
x-direction. Points Wy and By represent the y-motion of the wheel axis W and the body-suspension interaction point B, respectively. Lt
and bt is the compliance and damping and mass of the wheel tyre, respectively. Similarly, Ls and bs is the compliance and damping of

the wheel tyre, mw is the mass of the complete wheel and mb is the quarter of the mass of the car body. The model is excited by the
source of the velocity resulting from the tyre copying the road bumps.

1.7.2 Spring potential energy

From (1.30) and (1.31) we can see that the energy flow into an ideal rectilinear spring at a given instant is
(1.3
7)

If
, energy is being stored in the spring, and if
, the stored energy is being retrieved from the spring. Which of these two
states is occurring depends both on the spring deflection and relative velocity at the moment.
Integration of (1.37) gives an expression for the energy accumulated in the ideal rectilinear spring

which is called rectilinear potential energy.


We note that this energy depends directly on the force transmitted through the spring, but not on the velocity difference. Hence the
spring stores energy by virtue of its force. This energy is always positive and does not depend on the sign of either the force or the
spring deflection.

1.8 Summary
To set up a dynamic model of a real rectilinear system and to simulate its behavior using DYNAST, you should proceed through the
following steps.
1. Sketch the geometric configuration of your system, i.e.:
o clearly separate the system of your concern from the system surroundings and determine a reference frame for the
system
o associate the rectilinear motions with a coordinate axis
o for each body denote the geometric point representing its rectilinear motion and give it a name
o assume the positive velocity of each such point in the direction of the chosen coordinate axis
2. Set up a multipole diagram representing the dynamic model of your system and its surroundings (you may use directly the
DYNAST schematic editor):
o for each body place a node into the diagram corresponding to the geometric point representing the body velocity
o give each node a name identical with the name of the point (using a label in the DYNAST schematic editor)
o associate each of the nodes with the symbol for inertors

o place into the diagram one or more symbols for reference nodes representing the reference frame
o place symbols for elements representing couplings between the nodes (if necessary, interconnect the element pins with
the nodes by links)
o add symbols for elements representing the system surroundings
o invert the element symbols if their velocity is oriented in the opposite direction with respect to the chosen coordinate
axis
o give names to the inertors and the other elements and specify values of their parameters
3. Specify analysis of your system model:
o select the Numerical Nonlinear Analysis
o specify the required Time Interval of the analysis and set the values of Initial Conditions
o select Desired Variables like node and element velocities, element forces, and block outputs
o To invoke the analysis, select Run Analysis & Plot in menu Run
4. Interpret the analyzed results:
o Set the desired variables in menu Plot
o in menu Axes choose the way you want to plot the variables
o the assumed positive polarity of velocities are oriented in the direction of the chosen coordinate axes whereas the
assumed positive polarity of forces of the elements is opposite to this direction

Module 2
Rotational systems
Module units

2.1

Rotational interactions

2.1.1

Rotational motion

2.1.2

Torque

2.2

Ideal rotational inertors

2.3

Ideal rotational dampers

2.4

Ideal rotational springs

Module overview. In the preceding module we considered three mechanical elements representing three essential phenomena in
mechanical systems which undergo only translatory motions. Another large class of mechanical systems involves rotation about fixed

or nonaccelerating axes. Many types of rotating-machinery and control-system components, such as turbines, gear trains, electrical
machines, engines and pumps, may be modelled in part by the use of three pure rotary mechanical elements which are analogous to
the translational elements already discussed.
These rotary elements are rotational mass, or inertor; torsional or rotational spring; and torsional or rotational damper. The basic
physical phenomena involved in each of these pure physical elements is the same as for the corresponding translational element,
except that in rotary systems effects take place in terms of circular motions about an axis of rotation rather than in terms of straightline motion along an axis of translation. Also, torque, or moment of force, replaces force at points of interaction between system
elements.
Module objectives. When you have completed this module you should be able to:
1. Understand the fundamental laws governing dynamics of mechanical rotational systems.
2. Set up linear multipole models of these system dynamics from physical elements.
3. Use the computer package DYNAST to submit multipole diagrams representing the multipole models in the graphical form.
4. Use DYNAST to specify excitation and initial conditions of these models, invoke numerical time-domain analysis, and plot the
results.
Module prerequisites. This-course module on Translational mechanical systems.

2.1 Rotational interactions

2.1.1 Rotational motion


2.1.2 Torque

2.1.1 Rotational motion

In rotation about a fixed or nonaccelerating axis all points of the site other than those on the axis move in concentric circles about the
fixed axis.

Figure 2.1 Subsystems in rotational motion.


Fig. 2.1 shows a subsystem with two concentric shafts rotating about the z-axis fixed to the reference frame. The subsystem might be a
torsional spring, a rotational damper, or a gearbox, for example. The ends of the shafts A and B are in contact with shafts of some other
subsystems. The power intake of the subsystem due to the rotational motion of its energy entries A and B about the z axis is

where
and
are absolute angular velocities of the entries A and B about the z axis. A(t) and B(t) internal torques
counterbalancing external torques Aext(t) and Bext(t) applied to the entries.
Absolute angular motion about the axis z is described in Fig. 2.1 by the angular displacement , i.e., by the angle subtended
between a line drawn on the rotating entry which lies in the x-y-plane perpendicular to the axis of rotation z and the coordinate axis x

fixed to the reference frame. In this text, the counterclockwise direction will be considered as the positive orientation of rotation in a
plane perpendicular to the axis of rotation.
Absolute angular velocity

[rad/s] and absolute angular acceleration

[rad/s2] of an entry is by definition

where (t) is the angular displacement of the entry. The relative angular displacement, relative angular velocity and relative
angular acceleration of the entry B with respect to the entry A is

2.1.2 Torque

A moment of force or torque [N.m] can be viewed as the time rate of the angular momentum (t) acting in the same direction, i.e.

The complete specification of a moment of force (t) must include its magnitude and the axis of rotation. To have the energy flow
through an entry into a subsystem positive, the torques acting on the entry must be oriented in the opposite way with respect to the
direction of the entry angular velocity.

2.2 Ideal rotational inertors

An ideal rotational inertor models the rotational motion of a rigid body about an axis fixed to the reference frame. The constituency
relation of this physical element corresponds to Newtons second law which can be written as
(2.1
)

or as
(2.2
)

where (t) is the element rotational momentum, (t) is the element moment of force,
[kg.m2/rad] is the moment of inertia of the element.

is the element angular velocity, and J

Figure 2.2 (a) Symbol of rotational inertor. (b) Flywheel, and (c) its model.
Rotational inertors model inertial effect in rotational motion of bodies exclusively, they ignore any damping or spring effects as shown
in Fig. 2.2. The angular velocity
of a rotational inertor represents the absolute angular velocity of a body measured in relation to a
nonaccelerating reference frame. Thus the pole of rotational inertors must be always coalesced with a node corresponding to the
reference. This is the reason why only the +pole is shown in the element symbol.

If the moment of inertia JT of a body about a fixed axis passing through the mass center T is known, the moment of inertia J of the
body about another fixed axis parallel to the former one can be determined as
(2.3
)

where m is the mass of the body, and d is the separation of the axes.This formula, known as the parallel-axis or Steiners theorem, is
illustrated by Fig. 2.3. The diameter of the ball of the mathematical pendulum of Fig. 2.3a is so small that its moment of inertia is
negligible. If also the mass of the pendulum string is neglected, the pendulum moment of inertia with respect to the axis of rotation A
is J = mr2. Fig. 2.3b shows a compound pendulum. If only the moment of inertia JT with respect to mass center of the compound
pendulum body is known, the pendulum moment of inertia with respect to the axis of rotation A must be determined using (2.3). The
mass center of each of the pendulums is acted on by the moment of the gravitational force Fg = mg, the moment arm of length r sin
varies with the pendulum deflection . In the dynamic model for both pendulums shown in Fig. 2.3c, the ideal rotational damper of
factor B respects damping of the pendulums.

Figure 2.3 (a) Mathematical and (b) compound pendulum, (c) pendulum dynamic model.

2.3 Ideal rotational dampers

The constituency relation of an ideal rotational damper with the symbol of Fig. 2.4a is
(2.4
)

where (t) is the moment of force,


damper.

is the relative angular velocity, and br [N.m.s/rad] is the rotational damping constant of the

Figure 2.4 (a) Symbol of pure rotational damper. (b) Oil-filled rotational damper, and (c) its model.
An ideal rotational damper is a twopole model that ignores any mass or spring effects and assumes that a moment of force is
proportional to the relative angular velocity of its poles. An example of a real device with the - characteristic close to that of the ideal
damper, is the viscous rotational damper shown in Fig. 2.4b. It is a propeller rotating in oil inside a cylinder and producing the resisting
moment of force.
The poles represent -motions of the energy entries B and A of a real damping subsystem. represents the internal moment of force at
the entry associated with the +pole, i.e. at the entry B, counterbalancing an external moment of force applied to this entry. The positive
direction of is opposite to that for the damper angular velocity corresponding to the relative velocity of the entries
.
Since an ideal rotational damper has no moment of inertia, the moments of force acting on it must be balanced. Thus, an external
moment of force applied to Bx in the -direction flows through the damper and out through Ax without a change and, at the same time,

the internal moment of force at the damper pole denoted by Ax must be . Note that the power flow into the ideal damper
always positive regardless of the direction of its relative motion.

is

Fig. 2.5a shows the symbol of the ideal rotational damper used to model a shaft passing through a well lubricated journal bearing fixed
to the reference frame. In Fig. 2.5b, the same model is utilized for an oil-polluted clutch the discs of which are slipping with respect to
each other so that they rotate with different angular velocity and there is a viscous friction effect between them.

Figure 2.5 Ideal model of (a) a shaft in journal bearing, and (b) a slipping clutch.

2.4 Ideal rotational springs

For an ideal rotational or torsional spring, the constituency relation is


(2.5
)

or
(2.6
)

where (t) is the ideal spring angular displacement,


is the ideal spring angular velocity, (t) is the ideal spring moment of force, and
1
1
Lr [rad.N .m ] is the torsional compliance of the spring. The reciprocal value of Lr is the spring torsional stiffness or torsional spring
constant.

Figure 2.6 (a) Symbol of pure torsional spring. (b) Spiral torsional spring, and (c) its model. (d) Helical-coil spring. (e) Twisted rod.
The pure spring has no inertia by definition and hence transmits torque undiminished. One of the most common mechanical
components which exhibit the characteristics of a rotational spring is the torsional spiral spring shown in Fig. 2.6b. Fig. 2.6c shows a
helical-coil torsional spring, and Fig. 2.6d a twisted rod. BA corresponds to the relative angular deformation of the real spring, while
represents the spring internal moment of the force counterbalancing the external moment of a force applied to the spring.

Figure 2.7 (a), (b) torsional pendulum. (c), (d) double torsional pendulum.

Module 3
Coupled mechanical systems
Module units

3.1

Translatory-to-translatory couplings

3.1.1

Translational transformer

3.1.2

3.2

3.3

Levers

Rotary-to-rotary couplings

3.2.1

Rotational transformer

3.2.2

Gear trains

3.2.3

Gear backlash

3.2.4

Belt and chain systems

3.2.5

Belt friction

Translatory-to-rotary couplings

3.3.1

Translatory-rotational transformer

3.3.2

Rack-and-pinion

3.4

3.3.3

Pulley or sprocket assembly

3.3.4

Lead screws

3.3.5

Slider cranks

Load conversions

Module Overview. Thus far we have dealt with two classes of mechanical systems: those with all bodies in a translational motion
along the same straight line, and those with all bodies in a rotational motion about the same fixed axis. In both cases, the considered
mutual interactions of bodies were limited to dampers and springs.
In this module, we shall consider systems in which the bodies may translate in different directions, or rotate about different fixed axes.
The body motions in such systems are coupled by transducers of various forms. A transducer is defined as a device which transforms
energy or power in one form into energy or power in another form. A pure transducer performs this function with no energy loss or
storage, so that the net power flow into a pure transducer is zero. Here we shall introduce the concept of transduction by discussing
the pure mechanical transformer.
Common examples of transducers coupling bodies in different translational motion i.e. translatory-to-translatory transducers
represent levers and pulleys. Rotary-to-rotary couplings are associated with such components as gear trains, belts and pulleys,
sprockets and chains, or harmonic drives. Translatory-to-rotary transducers include rack-and-pinion, pulley-and-sprocket, lead-screw,
slider-crank, or cam assemblies.

Such mechanical transducers are used for the purposes like:

To change the speed of a motor or other power source to meet the need for a lower or higher output speed (as in the
automobile transmission which couples the engine to the driving wheels).

To obtain a nonuniform motion from a uniform one. This is a common application in automatic machinery.

To achieve a mechanical advantage, that is, an increase in torque level.

Module objectives. When you have completed this module you should be able to:
1. Understand the translatory-to-translatory, rotary-to-rotary, and translatory-to-rotary couplings in mechanical systems.
2. Set up both linear multipole models of these couplings using pure mechanical transformers.
3. Use the computer package DYNAST to submit multipole diagrams including the pure transformers in their graphical form.
Module prerequisites. Modules Translational mechanical systems and Rotational mechanical systems.

3.1 Translatory-to-translatory couplings

3.1.1 Translational transformer


3.1.2 Levers

3.1.1 Translational transformer

Translatory-to-translatory coupling is the coupling between the translational motion of one point and the translational motion of
another point along the same or different axis. Such couplings can be easily modelled by the pure translational transformer. It is an
idealized mechanical device which neither dissipates nor stores any energy and determines ratios of velocities and forces associated
with the two coupled motions. Fig. 3.1 shows our symbol for pure transformers. The variables associated in Fig. 3.1 with this symbol
correspond to the special case when the symbol is applied to a translational transformer.

Figure 3.1 Pure translational transformer.


Motion of point A with respect to point A_ along the x axis is represented by poles Ax and A_x, respectively. Similarly, motion of point
B with respect to point B_ along the axis is represented by poles A and A_. Fx is the force exerted by the transformer to move points
A and A_ apart while counterbalancing external forces applied to these points. F is the force counterbalancing external forces applied
to points B and B_.
The relation between velocities associated with the translational transformer is defined as

(3.1
)

where is the relative velocity of point A with respect to point A_, and is the relative velocity of point B with respect to point B_, n is
the transformer ratio. An ideal transformer has this ratio constant.
Forces associated with the translational transformer are related by definition as
(3.2
)

The power put into the translational transformer is


(3.3
)

After submitting (3.1) and (3.2), (3.3) becomes


. A pure translational transformer thus transforms a translational motion into
another translational motion and vice versa. It performs this function with no energy loss or storage, so that the net power flow into a
pure transformer is zero.
Table 3.1 shows examples of pure models of some translatory-to-translatory transducers. They include levers, pulleys and a rigid link
of a fixed length the endpoints of which are constrained to a linear or circular path.
Table 3.1 Examples of translatory-to-translatory motion converters.

Converter

Pure model

Coupling ratio n

3.1.2 Levers

A translatory-to-translatory coupling is typified by a lever which has no mass, is infinitely stiff, and is free from frictional effects. Such
a lever is shown in Fig. 3.2a.

Figure 3.2 (a) Lever, (b) its dynamic model.


The relationship between the velocities of endpoints A and B of the lever with respect to the pivot C is

(3.4
)

where n is the lever ratio. Note that a and b change as varies, but that a/b remains constant. Often will be limited to a small angle,
so that a and b are essentially constant.
The full-head arrows denote forces FA, FB and FC exerted by the lever at its end points A and B and at the pivot C, respectively, to
counterbalance external forces applied to these points. Since the massless lever has no momentum, then by definition the sum of the
forces in any direction and the moments of the forces about the pivot must be zero:
(3.5)
(3.6)

Therefore,
(3.7
)

Note, that the relationships between velocities (3.4) and between forces (3.7) are the same as in the translational transformer used in
Fig. 3.2b to model the lever.

3.2 Rotary-to-rotary couplings

3.2.1 Rotational transformer

3.2.2 Gear trains


3.2.3 Gear backlash
3.2.4 Belt and chain systems
3.2.5 Belt friction

3.2.1 Rotational transformer

Rotary-to-rotary coupling is the coupling between the rotational motion about a fixed axis and another rotational motion about the
same or different axis. Such couplings can be easily modelled by the pure rotational transformer. It is an idealized mechanical
device which determines ratios of angular velocities and torques associated with the coupled motions. Fig. 3.3 shows our symbol for
pure transformers. The variables associated with this symbol in Fig. 3.3 correspond to the special case when the symbol is applied to a
rotational transformer.

Figure 3.3 Pure rotational transformer.


Rotational motion of point A with respect to point A_ rotating about the same axis with angular velocity is represented by poles A
and A_, respectively. Similarly, the rotational motion of point B with respect to rotating point B_ with the velocity is represented by
poles A and A_. is the torque exerted by the transformer to move points A and A_ apart while counterbalancing external torques
applied to these points. is the force counterbalancing external torques applied to points B and B_.

Relation between angular velocities associated with the rotational transformer is


(3.8
)

where is the relative angular velocity of point A with respect to point A_, and is the relative angular velocity of point B with respect
to point B_. n is the transformer ratio. An ideal transformer has this ratio constant.
Torques associated with the rotational transformer are related as
(3.9
)

The power put into the rotational transformer is


(3.1
0)

After submitting (3.8) and (3.9), (3.10) becomes

A pure rotational transformer is thus an idealized mechanical device which transforms a rotational motion into another rotational
motion and vice versa. A pure transformer performs this function with no energy loss or storage, so that the net power flow into a pure
transformer is zero.

3.2.2 Gear trains

The simplest and most familiar example of a rotary-to-rotary coupling is the elementary gear train shown in Fig. 3.4a. Two mating
gears of different sizes are attached to two parallel shafts which in turn are free to rotate supported in a housing. Gear-tooth profiles
machined into the periphery of the gear disks cause the gears to behave like circular cylinders of radii ra and rb, which roll on each
other without slipping, provided that the gears are rigid and the gear teeth are geometrically perfect. The radii ra and rb are called the
pitch radii of the gears, and their magnitudes are proportional to the numbers of teeth Na and Nb. The gear ratio is defined as

Figure 3.4 (a) Gear train, (b) its free-body diagram, (c) transformer model.
Since the gears are assumed to roll without slipping, the arc distance
turned off on gear b. Hence

turned off on gear a must equal minus the arc distance

(3.1
1)

where the negative sign indicates that the direction of

is opposite to that of

To determine the relationship between the torques A and B, let us isolate the two gears from each other and from the frame, thus
forming their free-body diagrams of Fig. 3.4b. Only torques and torque-producing forces are shown on these free bodies, and friction

between the gear-tooth surfaces is assumed to be zero. If we further assume that the gears have zero inertia or constant speed, the net
torque on each gear is zero. If we also assume the bearing torques to be negligible,

Employing now the definition of n in (3.11), we obtain


(3.1
2)

Thus the speed is changed as n while the torque is changed as 1/n. The net power into the gear train for

is therefore zero:

Since the gear system is assumed to have constant angular momentum, the torque C shown acting on the container Fig. 3.4a must be
equal to minus the sum of A and B, or
(3.1
3)

Since

in the preceding discussion, C does no work on the system.

Table 3.2 shows various setups of gear trains that can be modeled by a single pure rotational transformer. The table gives the
corresponding gear ratios.
Table 3.2 Rotary-to-rotary motion converters

Gear train
External
spur gears

Internal
spur gears

Bevel
gear pair

Planet
gear

Skew
gear pair

3.2.3 Gear backlash

Configuration

Coupling ratio n

Fig. 3.5 indicates the way in which a pure model of a pair of spur gears can be augmented to take into the consideration backlash
between the gears. Fig. 3.5b shows a simple characteristic approximating the torque exerted by the gears on each other in relation to
the angle AB between the gears. The finite slope of the nonzero branches of the characteristic corresponds to the torsional stiffness of
the gear teeth. The backlash effect is implemented in the model by adding the source of torque controlled by the angle AB in
accordance with the characteristic shown in Fig. 3.5b to the energy-transfer transformer modeling the backslashes motion conversion.
Also slippage in rotary-to-rotary motion converter due to friction can be modelled in a similar way.

Figure 3.5 (a) Gear train with backslash, (b) backslash characteristic, (c) multipole model.

3.2.4 Belt and chain systems

Figure 3.6 Belt-and-pulley and chain-and-sprocket assemblies.


Fig. 3.6 shows pure belt-and-pulley and chain-and-sprocket systems that perform essentially the same function as gear trains. As
opposed to a belt and pulley, the latter system does not allow slippage between the belt and sprocket due to the sprocket teeth. Note
that the sign of n is positive in the case of the system in Fig. 3.6a as the pulleys rotate in the same direction, whereas in the case of the
system in Fig. 3.6b the sign is negative.

3.2.5 Belt friction

Fig. 3.7a shows a flat belt passing over a fixed cylindrical drum. To model belt slipping let us consider interactions between segments
of the belt and drum surface which are in contact. This can be taken as a one-dimensional translational problem if the belt segment as
well as the adhering drum surface is straighten up as shown in Fig. 3.7b. The variables can be then related to the coordinate axis x.

Figure 3.7 Belt friction.


The Eytelweins formula gives the relation between the forces FA and FB acting at the belt-segment endpoints A and B. For the case
when the belt is just about to slip if pulled to the right, i.e. if FA > FB > 0, the formula is:
(3.1
4)

where [] is the coefficient of friction, and [rad] is the angle of the belt contact. This formula is derived under the assumption that
the belt segment is massless and rigid, and the friction between the belt and the drum is governed by the law of dry friction.
A multipole diagram of the belt friction constructed under the above assumptions is given in Fig. 3.7c. The contact areas of the belt and
drum segments are represented there by poles A and C, respectively. Friction is modelled by a source of friction force Ff acting
between these two poles. The rigid belt segment is modelled by the source of zero velocity
placed between the poles A and B

representing the belt-segment end points. At the same time, this source is employed in the model as a sensor of the tension in the belt
segment. This tension force Ft is used to control the friction-force source Ff.
Taking into consideration the balance relations for forces at the poles
(3.1
5)

and substituting them into the formula we shall obtain the constitutive relation of the friction-force source
(3.1
6)

The function sgn(Ft) was included into the formula to make it valid also for the case when the belt is pulled to the left, i.e. for FB > FA
> 0. (You are encouraged to check its correctness.)
Equation (3.16) can be used both for static friction of impending slipping as well as for the slipping kinetic friction after substituting
there either the coefficient of static friction s or the coefficient of kinetic friction k for . Fig. 3.8 shows an example of the Ffcharacteristic of combined static and kinetic friction governed by (3.16) for the tension
. Note that the characteristic is
asymmetric with respect to the origin.

Figure 3.8 Belt friction characteristic.


As shown in Fig. 3.7d, the belt friction model from Fig. 3.7c can be completed to respect the belt segment flexibility kb, and mass mb.
The belt friction model we have derived for moving belt and fixed drum applies equally well to problems involving fixed belt and
rotating drum, like band brakes, and to problems involving both belt and drum rotating, like belt drives. It can be also used to systems
with the belt replaced by a rope or a string.
In case of V-shaped belts (3.16) should be replaced by
(3.1
7)

where is the angle of the V of the belt cross section.


Fig. 3.9a shows an example of a hoisting machine lifting a load by means of a cable passing over an undriven sheave and wrapped
around a driven drum. The rotary motion of the drum drive is transformed into the translational motion of the cable by means of an
ideal transformer with the ratio 1 : r1, where r1 is the drum radius. The c1 and c2 parts of the cable are modelled by elements
characterizing the cable flexibility and damping. The element parameters of the cable part c2 depend on the cable length of c2. The
friction between the cable and the sheave is represented by the belt-friction model given above. The sheave dynamics, converted from
rotary to translational motion along the cable, is represented by the elements of reduced parameters
(3.1
8)

where Js is the moment of inertia of the sheave, r2 is its radius and bs is the damping of the sheave bearing.

Figure 3.9 Hoisting machine.

3.3 Translatory-to-rotary couplings

3.3.1 Translatory-rotational transformer


3.3.2 Rack-and-pinion

3.3.3 Pulley or sprocket assembly


3.3.4 Lead screws
3.3.5 Slider cranks

3.3.1 Translatory-rotational transformer

Translatory-to-rotary coupling is a coupling between the translational motion and a rotational motion about a fixed axis. Such
couplings can be easily modeled by the pure translatory-rotational transformer. It is an idealized mechanical device which
determines ratio of the velocities as well as the ratio of the force and torque associated with the two coupled motions. Fig. 3.10 shows
our symbol for pure transformers. The variables associated with this symbol in Fig. 3.10 correspond to the special case when the
symbol is applied to a rotational transformer.

Figure 3.10 Pure translatory-rotational transformer.


Rotational motion of point A with respect to point A_ along the x axis is represented by poles Ax and A_x, respectively. Rotational
motion of point B with respect to point B_ rotating about the same axis with the velocity is represented by poles A and A_. F is the
force exerted by the transformer to move points A and A_ apart while counterbalancing external forces applied to these points. is the
torque counterbalancing external torques applied to points B and B_.
Relation between angular velocities associated with the rotational transformer is

(3.1
9)

where is the relative velocity of point A with respect to point A_, and is the relative angular velocity of point B with respect to point
B_. n is the transformer ratio. An ideal transformer has this ratio constant.
The force and torque associated with the translatory-rotary transformer are related as
(3.2
0)

The power put into the rotational transformer is


(3.2
1)

After submitting (3.19) and (3.20), (3.21) becomes


. A pure translatory-rotary transformer is thus an idealized mechanical device
which transforms a translational motion into a rotational motion and vice versa. A pure transformer performs this function with no
energy loss or storage, so that the net power flow into a pure transformer is zero.

3.3.2 Rack-and-pinion

In motion control problems it is often necessary to convert rotational motion into a translational one and vice versa. For instance, a
load may be controlled to move along a straight line through a rotary motor and rack-and-pinion assembly, such as that shown in Fig.

3.11a. The pinion is a small gear, and the rack is a linear member with the gear on one side. The describing relation of this system is
defined as
(3.2
2)

Figure 3.11 (a) Rack-and-pinion gear train, (b) equivalent system, (c) pure model.
Thus the constitutive relation of pure rotary-linear motion converters is

where the coupling ratio

Figure 3.12 Movable rack-and-pinion assembly.


To apply this result to the rack-and-pinion assembly movable along the x-axis as that shown in Fig. 3.12a, the variable x should be
replaced by xAB.

3.3.3 Pulley or sprocket assembly

Another approach to rotary-translatory motion conversion is to utilize the rotary motion of a pulley combined with the linear motion of
a belt. The same effect is achieved by combining a sprocket with a chain, a drum with a rope or cable wound around it, etc. This
concept is outlined in Fig. 3.13a.
A common assembly of two pulleys and a belt used for the control of a mass through one of the pulleys by a rotary prime mover is
shown in Fig. 3.13b. Both the drive and idle pulley have the same radius r.

Figure 3.13 (a) Pulley or sprocket. (b) Belt and pulley rotary-translatory assembly.

3.3.4 Lead screws

Fig. 3.14 shows a similar situation in which a lead screw is used as the rotary-to-translatory mechanical linkage. The lead screw is used
there to drive a payload along the z-axis. The screw is fixed free to rotate and as it is turned, the nut moves along the screw with the
payload attached. In this case, the describing relation is defined as the screw rotation per unit of linear motion x. This is also referred
to as the pitch P of the screw (the reciprocal of the pitch is called the lead). Therefore, the describing relation is

and the coupling ratio of the lead screw is simply

The sign is related to the mutual orientation of the rotational and translational motions.

Figure 3.14 Lead screw assembly.

3.3.5 Slider cranks

Also various mechanical linkages are used as rotary-translatory converters. As an example of this approach, Fig. 3.15 shows a
configuration of a crank driving a linear stage. The crank of length r rotates about its center A and has a rod of fixed length l mounted
to its rotating end. The other end of the connecting rod B is attached to a linear stage which is constrained to move in only along the xaxis. As the describing relation is

the reciprocal of the coupling ratio is

Figure 3.15 Slider crank.


As the crank rotates by
, the linear stage moves in both directions along x. There are two distinct angular positions of the crank
that correspond to full extension and retraction of the stage called dead-center positions. If this module is used to convert linear to
rotary motion the direction of rotation may go in either way in these positions (this problem is usually overcome by combining the
crank with a flywheel).

3.4 Load conversions

Figure 3.16 (a) Cascade interconnection of energy-transfer transformers, (b) one-transformer equivalent.
Fig. 3.16a shows a cascade interconnection of energy-transfer transformers of coupling ratios n1 and n2. A simple analysis of the total
transformation of across and through variables in the interconnection indicates that the cascade can be replaced by only one
transformer with the coupling ratio n1n2 shown in Fig. 3.16b.

Figure 3.17 Load seen through an energy-transfer transformer.


Fig. 3.17 indicates the way in which series or parallel combinations of physical elements connected to a pole-pair of an energy-transfer
transformer are seen through the transformer, or referred to at the other transformer pole-pair. Z resp. Y is an impedance resp.
admittance of the generic element, E resp. J is an independent source of an across resp. through variable. This is the principle behind
utilizing various transducers and couplings to convert the amount of an actuator load at the price of changed related across and through
variables.

Module 4
Nonlinear mechanical systems
Module units

4.1

Fluid friction

4.1.1

4.2

Dry friction

4.3

Updated friction models

4.4

Belt friction

Gear backlash.

Module overview. This module will introduce you into dynamic behavior of nonlinear or time-variable mechanical systems the bodies
of which are in a translational rectilinear or rotational motion.
Module objectives. When you have completed this module you should be able to:
1. Understand basic principles of dynamics of nonlinear mechanical systems.
2. Set up multipole diagrams representing dynamic models of these systems from physical elements.
3. Use the computer package DYNAST to submit multipole diagrams representing the multipole diagrams in their graphical form.
4. Use DYNAST to specify excitation and initial conditions of these models, invoke numerical time-domain analysis, and plot the
results.
Module prerequisites. Modules Translational mechanical systems Rotational mechanical systems

Jumping ball 1

System
A ball falls down on a rigid surface.

To see animations,
please get Adobe Flash

Sites of Interaction
B

...

ball mass center

System excitation
The initial altitude of the ball above the surface is 0.2 m.

Task
Plot the vertical diplacement of the ball above the surface.

Assumptions
The ball compliance and also its damping is constant during its contact with the surface.

System Parameters
m=2

[kg]

ball mass

c = 1

[m/N]

ball compliance

d = 100

[N.s/m]

ball damping factor

g = 9.81

[m/s2]

gravitational constant

r = 0.05

[m]

ball radius

h0 = 0.5

[m]

initial height

tf = 5

[s]

final time

Solution
Model

The time-variable contact between the ball and the surface is modeled by the switch which becomes closed each time the ball touches
the surface.
Results

Data

The following data were generated by DYNAST for the above given system model. You can modify the model parameters and resubmit the data to DYNAST to receive new results.

:jumball1
*: Jumping ball 1
*SYSTEM;
m = 2;
::[kg] ball mass
c = 1u; ::[m/N] ball compliance
d = 100; ::[N.s/m] ball damping factor
g = 9.81; ::[m/s^2] gravitational constant
r = 0.05; ::[m] ball radius
h0 = 0.5; ::[m] initial height
tf = 5;
::[s] final time
:---------------------------------------------------S1 S-B = yB < r; :sw itch closes w hen yB < r

Show results as

Submodels

Integrator block

Origin

H. Mann
Mass-spring system with dry friction

Purpose
Modeling of dry friction, usingn trimmed linear function

System
A body resting on a frictional surface is displaced from its steady-state position by a spring.

To see animations,
please get Adobe Flash

Sites of Interaction
B

...

body

System excitation
The body is given an initial dicplacement xB(0) = 4 m.

Task
Analyze the time variation of the block displacement xA, velocity

and acceleration

Assumptions
Dry friction between the body and the surface can be approximated by the restraining force characteristic shown in Fig.(b).

System Parameters
m=1

[kg]

block mass

k=1

[N/m]

spring stiffness

R0 = 1

[N]

friction force amplitude

k0 = 10k

[N/m]

local stiffness of dry-friction approximation

tf = 8

[s]

final time

xB0 = 4

Solution
Model

Modeling of the spring by the source of the spring restoring force SPRING allows for specifying the spring initial displacement. The
force-velocity characteristic of the dry friction approximation is specified by a linear function of slope k0 trimmed at
and with
slope set to zero. The local stiffness k0 avoids a step discontinuity as the velocity changes sign and allows the numerical integration to
proceed smoothly. The stiffness k0 represents the local stiffness of the surfaces in contact.

Results

As shown, for example, in [1], the governing equation for the system is

where = k/R0. The upper sign is used when the velocity is positive, and the lower sign is used when it is negative. The natural
frequency of the system
is not changed by the frictional efect. The motion is sinusoidal about the point
,
depending on the velocity in that swing. The end of the first swing occurs at a time t = /, so that displacement is xB = xB(0) + 2.
During the second swing, the sinusoid is centered about xB = and the displacement at the end of the swing is xB = xB(0) 4. It is
seen that the displacement at the end of each swing is reduced by an amount of 2.
Observe that the numerical results are indeed those predicted by theory. The motion is seen to cease at the end of the second swing,
which is also predictable through the above given equation.

Data

The following data were generated by DYNAST for the above given system model. You can modify the model parameters and resubmit the data to DYNAST to receive new results.

*: Mass-spring system w ith dry friction


*SYSTEM;
m = 1;
::[kg] block mass
k = 1;
::[N/m] spring stiffness
R0 = 1; ::[N] friction force amplitude
k0 = 10k; ::[N/m] local stiffness of dry-friction approximation
tf=8;
:: [s] final time
:--------------------------------------------------xB0=4;
:Dry friction characteristic:
dryf /LIN/ B=k0, L=-R0/k0, U=R0/k0, SL=0, SU=0;
R > J Bx = dryf(Bx);
:dry-friction force

Show results as

Submodels

Integrator block

Origin

H. Mann
[1] Levy S., Wilkinson J.P.D.: The Component Element Method in Dynamics. McGraw-Hill 1976, p. 50.
[2] Mann H.: Computer-Aided Design of Dynamic Systems (in Czech). House of Engineering, Prague 1986.

Motor on vibration isolator with stop

Purpose
Modeling of a nonlinear spring and repeated transient analysis.

System
A motor is mounted on a vibrational isolator consisting of a linear spring and a damper. A stop is added to restrain the motor motion
excited by an unbalanced force.

To see animations,
please get Adobe Flash

Sites of Interaction
A

...

motor body

System excitation
Motor is acted on by a sinusoidal imbalance force F = 44.5 sin 25t N. The motor is assumed to be initially at rest, i.e., yA(0) = 0.

Task
Study the effect of the stop clearance on the acceleration of the motor for yS = 12 mm, yS = 8 mm, and yS = 2.5 mm in the time interval
s.

Assumptions
The stop restoring force characteristic can be approximated as shown in Fig.(b). The gravitational force acting on the motor can be
neglected.

System Parameters
m = 17.5

[kg]

motor mass

k = 11k

[N/m]

spring stiffness

d = 175

[N.s/m]

damping

kS = 110k

[N/m]

stop stiffness

yS = 2.5m

[m]

stop clearance

tf = 0.8

[s]

final time

Solution
Model

Results

We observe that for a clearance of 12 mm no contact is made with the stop. However, when the clearance is reduced to 8 mm, contact
is made by the third oscillation and a large acceleration occurs. With a smaller gap of 2.5 mm, similar contact accelerations occur even

sooner. The acceleration could be reduced by decresing the stop stifness kS, but in that case, of course, the displacements would
increase.
Data

The following data were generated by DYNAST for the above given system model. You can modify the model parameters and resubmit the data to DYNAST to receive new results.
: File TRA9 Date 08/04/1999 Time 17:32:47
*: Motor on vibration isolator w ith stop
*SYSTEM;
:parameters:
m = 17.5;
::[kg] motor mass
k = 11k;
::[N/m] spring stiffness
d = 175;
::[N.s/m] damping
kS = 110k;
::[N/m] stop stiffness
yS = 2.5m;
::[m] stop clearance
tf = 0.8;
::[s] final time
:----------------------------------------------:motor:

Show results as

Submodels

Integrator block

Origin

H. Mann

[1] Levy S., Wilkinson J.P.D.: The Component Element Method in Dynamics. McGraw-Hill 1976, p.39.
[2] H. Mann: Computer-Aided Design of Dynamic Systems (in Czech). House of Engineering, Prague 1986.

Transformer mounted in a case*

Purpose
Modeling of an impact and frequency-variable excitation.

System
Equipment package contains a transformer suspended from the outer case by a spring. The transformer is free to move through a
clearance before contacting spring-stops fixed to the case. The case is placed on a shake table.

To see animations,
please get Adobe Flash

Sites of Interaction
T

...

transformer body

...

outer case and shake table

System excitation
The excitation is that of the shake table driven at 10 g over a linear sweep of frequency from 20 to 60 Hz in 1 s. The transformer is
initially at rest.

Task
We want to observe the growth of the acceleration amplitude sustained by the transformer during the frequency sweep.

Assumptions
Restoring force characteristic can be considered as piece-wise linear as shown in Fig. . The transformer weight need not be considered,
as the system configuration specification applies to the steady state.

System Parameters
m = 9.07

[kg]

transformer mass

d = 263

[N.s/m]

damping factor

k = 525k

[N/m]

spring stiffness

kS = 2.63me

[N/m]

stop stiffness

cS = 5m

[m]

stop clearance

g = 9.81

[m/s2]

gravitational acceleration

tf = 0.8

[s]

final time

Solution
Model

The time-variable frequency of the excitation is given as = 2(20 + 50t) rad/s. Thus the table acceleration ay can be computed as ay =
10g sin , where is obtained by integrating 2. The source of force modeling the shake-table is controlled by the acceleration
integrated.

Results

The results are plotted with a part of the 1s period omitted to save space. We note that the upper stop is contacted first at about 0.315 s
when the stop clearance is cS = 5 mm. In the next half-cycle, the lower stop is contacted. Continued periodic contact with the stops

occurs as the frequency is increased to 60 Hz. The maximum acceleration reaches 110 g during the contact. The steady-state amplitude
would have been only about 17 g if there had been no stops, or if the stops had not been contacted in going to 60 Hz.
When the stop clearance is cS = 1 mm, the first stop contact occurs during the first period already.
Data

The following data were generated by DYNAST for the above given system model. You can modify the model parameters and resubmit the data to DYNAST to receive new results.
*: Transformer mounted in a case*
*SYSTEM;
m = 9.07; ::[kg] transformer mass
d = 263; ::[N.s/m] damping factor
k = 525k; ::[N/m] spring stiffness
kS = 2.63me; ::[N/m] stop stiffness
cS = 5m; ::[m] stop clearance
g = 9.81; ::[m/s^2] gravitational acceleration
tf = 0.8; ::[s] final time
:----------------------------------------------------:model:
omega = 20 + 40*time;
:shake-table variable frequency

Show results as

Origin

H. Mann
[1] Levy S., Wilkinson J.P.D.: The Component Element Method in Dynamics. McGraw-Hill 1976, pp. 4749.

[2] H. Mann: Computer-Aided Design of Dynamic Systems (in Czech). House of Engineering, Prague 1986.

Rocket liftoff*

Purpose
Modeling of time-variable mass and variable gravitation.

System
The first stage of Saturn V rocket is fired vertically up from the earths surface at the north pole. It accelerates until the fuel, which
burns at a constant rate, is exhausted.

To see animations,
please get Adobe Flash

Task
Compute rocket velocity, position, mass, thrust and the gravitational acceleration within the first 300 s of the liftoff.

Assumptions
The back pressure across the nozzle and the aerodynamic resistance can be neglected as the velocity of the ascending rocket is smallest
in the dense part of the atmosphere. The relative nozzle velocity of the exhaust gas has a constant value, and the nozzle exhausts at
atmospheric pressure throughout the flight. As the liftoff takes place at a pole any influence due to the earths rotation can be
eliminated.

System Parameters
mr = 6.32me

[kg]

mass of rocket without fuel

mf = 4.56me

[kg]

launch mass of fuel

t1 = 160

[s]

period of fuel burnout

u = 8470

[m/s]

exhaust velocity of burned gases

r = 6.371me

[m]

radius of earth

g0 = 9.83

[m/s2]

gravitational acceleration at poles

tf = 300

[s]

final time

Solution
Model

The rocket of variable mass m(t) can be considered as one system together with the ejected fuel of mass m0(t) the total mass m + m0 of
which remains constant. According to the generalized Newtons second law, the total external forces F acting on such a system must be
in dynamic equilibrium with the time rate of changes of the total linear momentum of the system mass, i.e.

where

is the absolute velocity of the rocket and v0 is the absolute velocity of the ejected fuel.

Note that
and that
since m0 moves undisturbed with no acceleration once free of the nozzle. Thus the total external
forces F on the system must be in equilibrium with

where
is the thrust of the rocket, and u = v v0 is the relative velocity of exhaust gasses with respect to the rocket. Therefore,
the rocket dynamics may be modeled by a time-variable inertor to which a source of the thrust T is applied.
It may be observed that for a given mass rate of flow, the rocket thrust T depends only on the relative exhaust velocity u. It is
independent of the magnitude and of the direction of the absolute velocity v0 of the exhaust gases. (During the initial stages of motion
when v < u, v0 will be directed rearward, whereas when the rocket reaches a velocity v > u, v0 will be directed forward.)

Variation of the gravitational acceleration g is determined by the gravitational law as

Results

Data

The following data were generated by DYNAST for the above given system model. You can modify the model parameters and resubmit the data to DYNAST to receive new results.
*: Rocket liftoff*
*SYSTEM;
mr = 6.32me; ::[kg] mass of rocket w ithout fuel
mf = 4.56me; ::[kg] launch mass of fuel
t1 = 160;
::[s] period of fuel burnout
dotm = -mf/t1; :mass rate of burning fuel
u = 8470;
::[m/s] exhaust velocity of burned gases
r = 6.371me; ::[m] radius of earth
g0 = 9.83; ::[m/s^2] gravitational acceleration at poles
tf = 300;
::[s] final time
:---------------------------------------------------------m = mr

Show results as

Submodels

Integrator block

Origin

H. Mann
[1] H. Mann: Computer-Aided Design of Dynamic Systems (in Czech). House of Engineering, Prague 1986.
[2] Meriam J.L., Kraige L.G.: Engineering Mechanics. Dynamics. Wiley 1998, pp. 8, 309310.

4.1 Fluid friction

Mechanical resistances are due to the effects of friction caused by the physical interactions between body surfaces moving relatively to
each other. Friction is a key characteristic of gearings, screws and seals, as well as of gears, hydraulic components, electromechanical
transducers, etc. Friction is present to some degree in all mechanical systems. In many of them, friction consumes a major portion of
the applied energy.
Let us consider the experiment shown Fig. 4.1a where a body is placed on a horizontal plane surface. If there are no external forces
acting on the body besides the vertical force Fn (including the body weight), the reaction of the surface Fr = Fn is normal to the contact
plane. When an external horizontal force F is applied to the body, a reaction force component Ff tangential to the contact plane and
opposing F is developed. Ff is called the friction force as its origin is due to the friction between the body and the surface in contact.

Figure 4.1 (a) Raise of sliding friction, (b) port-diagram model, (b) fluid friction characteristic.
Fig. 4.1b shows a port-diagram model of the experiment. The effect of friction is represented there by a pure damper. The pair of port
variables associated with the damper consists of the friction force Ff and of the relative velocity of the body and the surface.
Friction between well lubricated surfaces can be usually approximated by the friction model called fluid or viscous friction. Its
characteristic Ff , given in Fig. 4.1c, shows that in this friction model it is assumed that the friction force Ff is a linear function of
the relative velocity between the contact surfaces, independent of the normal force Fn. The slope b of this function is called the

damping factor of the fluid friction. In this case the effect of friction can be respected in the port-diagram model by an ideal damper
with a constant resistance b.

4.1.1 Gear backlash.

Fig. 3.5 indicates the way in which a pure model of a pair of spur gears can be augmented to take into the consideration backlash
between the gears. Fig. 3.5b shows a simple characteristic approximating the torque exerted by the gears on each other in relation to
the angle AB between the gears. The finite slope of the nonzero branches of the characteristic corresponds to the torsional stiffness of
the gear teeth. The backlash effect is implemented in the model by adding the source of torque controlled by the angle AB in
accordance with the characteristic shown in Fig. 3.5b to the energy-transfer transformer modeling the backlashless motion conversion.
Also slippage in rotary-to-rotary motion converter due to friction can be modeled in a similar way.

Figure 4.2 (a) Gear train with backslash, (b) backslash characteristic, (c) multipole model.

4.2 Dry friction

Most of the textbooks on engineering mechanics recognize just two types of friction opposing relative motion of bodies in mechanical
contact: fluid or viscous friction, and dry or Coulomb friction. We have seen in modules on translational and rotational systems already
that modeling of fluid friction of a constant damping factor is quite easy.
Dry friction, however, is often neglected just because it is poorly understood and considered as difficult to model. Yet dry friction may
effect machine behavior in a way deteriorating or even disabling machine control. Dry friction-induced vibrations, often referred to as
stick-slip, are causing such industry plaguing effects like break and bearing squeals or chatter of machine tools. Dry friction modeling
is not an unattainable goal, however.
Dry friction between two rigid bodies in mechanical contact is commonly assumed to be caused by a dry friction force of two
components - the kinetic friction force and the static friction force. Both these reaction forces are developing in the direction tangential
to the plane of contact and both are opposing the relative motion of the contact surfaces.
The kinetic friction force Ffk can be specified by the expression
(4.1)
where k is a constant called the coefficient of kinetic friction, Fn is the normal force driving the two surfaces into contact, and is the
relative velocity of the two surfaces. The signum function is defined as

(4.2)

At zero relative velocity , it is the static friction force that tends to prevent motion. It is usually defined as
(4.3)

where s is the constant coefficient of static friction. In most cases, s > k.


The two dimensionless coefficients of friction can thus be defined as
(4.4)
Values of both the coefficients are assumed to be independent of the geometric area of the mechanical contact.
Fig. 4.3 displays two different characteristics of the dry friction force defined above. In Fig. 4.3a the friction force Ff is plotted with
respect to the relative velocity of the contact surfaces whereas Fig. 4.3b shows the way in which the dry friction force Ff is related to
the tangential force F applied to the contact surfaces. It is important to note in this figure that a part of the displayed characteristic for
increasing F differs from that for decreasing F.

Figure 4.3 Dry friction characteristics.


From the characteristics in Fig. 4.3 we can observe that three different situations may occur when two surfaces are in mechanical
contact:
1. The surfaces are in mutual rest and the applied force F is smaller (in absolute value) than Fs. F is thus not large enough to set
the surfaces in motion as the surface reaction, the friction force Ff, develops to be in equilibrium with F. The surfaces remain in

rest up to the point when the applied force F reaches the value Fs. Then the surfaces are just about to slide along each other
their motion is impending.
2. The applied force F has exceeded the value Fs. Then the friction force Ff cannot balance the applied force F any more and the
surfaces start sliding. As soon as they were set into motion, the magnitude of Ff drops into a lower value Fk. From there on the
surfaces keep sliding while the friction force remains constant at the value Ff = Fk as far as the force F provided that it does not
drop below the value Fk.
3. The applied force F has dropped below the value Fk. The friction force Ff gets into balance with F and the sliding motion stops.
As the dry friction description given above was formulated rather early (Amontos in 1699, Coulomb in 1755) it has been termed as the
law of dry friction, and the coefficients of friction are treated as intrinsic material constants. Despite this, the expressions (4.1) and
(4.3) should be regarded only as a specification of a rough empirical model of limited validity.
And more than that, this model is inconvenient for numerical computations because of its discontinuities at zero velocity. Besides
functions specifying any nonlinearities, the algorithms for numerical integration of differential equations operate also with the
derivatives of these functions. Most of the algorithms will just collapse when encountering a characteristic of dry friction given above.
The robust algorithms will have to shorten the integration steplength to such an extent that the computations will slow down very
significantly.

4.3 Updated friction models

According to tribological studies, two solid surfaces in contact over a certain geometric area are actually touching only in discrete
spots microcontacts. The macroscopic frictional force is thus the resultant of a great number of microscopic friction forces acting at
the individual microcontacts due to their elastic and plastic deformation, adhesion bonding and other effects.
Careful experiments reveal, that there are no discontinuities in the measured friction-force characteristics in reality. Besides that it was
found, that near zero velocity, friction is a continuous function of displacement.

In many practical applications, it is quite sufficient to approximate the kinetic friction by the piece-wise linear characteristic shown in
Fig. 4.4a. This approximation is continuous yet simple, so its evaluation requires few computer operations.

Figure 4.4 Continuous approximations to the dry friction characteristic.


For both kinetic and static friction the exponential approximation shown in Fig. 4.4b was proposed. It is specified as
(4.5
)

where a and b are parameters related to surface roughness or the bonding ability of the surfaces and Fr is the ratio of the asymptotic
value of the curve to its maximum. In a sense, Fr gives the ratio of the static and kinetic friction forces. The parameter b is related to
the static, and the parameter a to the kinetic friction components.
In this model, also the first derivative of the approximation is continuous (with the exception of the zero velocity). Most of the friction
force expressions utilized in the literature can be recovered from Eq. (4.5) by appropriate combinations of parameters a, b and Fr.
Fig. 4.5 shows a simplified Stribeck curve representing the general characteristic of lubricated moving surfaces as a function of the
lubricant viscosity , the relative velocity , and the normal force Fn.

Figure 4.5 Stribeck curve.


Depending on the geometry, the materials, the operating conditions and the separation of the contacting surfaces three main lubrication
regimes may be distinguished. In regime I, the rigid surfaces are separated by continuous lubricant film, whose thickness is much
larger than the surface roughness. The friction is due to to the internal friction of the lubricant. In this regime, the tribological behavior
of the system is determined by the rheology of the lubricant and can be calculated by the methods of fluid mechanics.
The published theoretical and empirical parameter values of the various friction models published in the literature provide only rough
estimates of the behavior in a particular system, however. While it may be possible to use them to identify the dominant sources and
effects of friction in a system, the actual parameter values should be identified by experiment.

4.4 Belt friction

Fig. 4.6a shows a flat belt passing over a fixed cylindrical drum. To find a model of belt slipping with respect to the drum let us
consider the interactions between the segments of the belt and of the drum surface which are in contact. This can be taken as a onedimensional translational problem if the belt segment as well as the adhering drum surface is straighten up as shown in Fig. 4.6b. The
variables can then be related to the coordinate axis x.

Figure 4.6 Belt friction.


In textbooks on engineering mechanics one can find Eytelweins formula giving the relation between the forces FA and FB acting at the
belt-segment endpoints A and B. For the case when the belt is just about to slip if pulled to the right, i.e. if FA > FB > 0, the formula is:
(4.6
)

where [] is the coefficient of friction, and [rad] is the angle of belt contact. This formula is derived under the assumption that the
belt segment is massless and rigid, and that the friction between the belt and the drum is governed by the law of dry friction.
A port-diagram model of the belt friction constructed under the above assumptions is given in Fig. 4.6c. The contact areas of the belt
segment and of the drum are represented there by the poles A and C, respectively. The friction is modeled there by a source of friction
force Ff acting between these two poles. The rigid belt segment is modeled by the source of zero velocity
placed between the poles
A and B representing the belt-segment end points. At the same time, this source is employed in the model as a sensor of the tension in
the belt segment. This tension force Ft is used to control the friction-force source Ff.
Taking into consideration the balance relations for forces at the poles
(4.7
)

and substituting them into the formula (4.6) we shall obtain the constitutive relation of the friction-force source
(4.8
)

The function sgn(Ft) was included into the formula to make it valid also for the case when the belt is pulled to the left, i.e. for FB > FA
> 0. (You are encouraged to check its corretness.)
Eq. (4.8) can be used both for static friction of impending slipping as well as for the slipping kinetic friction after substituting there
either the coefficient of static friction s or the coefficient of kinetic friction k for . Fig. 4.7a shows an example of the Ff
characteristic of combined static and kinetic friction governed by Eq. (4.8) for the tension
. Note that the characteristic is
asymmetric with respect to the origin.

Again, for practical computations it is recommended to replace the discontinues characteristic based on the law of dry friction by a
more conveniently shaped approximation function like those shown in Fig. 4.4 (see Fig. 4.7b). These functions have to be multiplied
then by the right-hand side of Eq. (4.8).

Figure 4.7 Belt friction characteristic.


As shown in Fig. 4.6d, the belt friction model from Fig. 4.6c can be completed to respect the belt segment flexibility kb, and mass mb.
The belt friction model we have derived for moving belt and fixed drum applies equally well to problems involving fixed belt and
rotating drum, like band brakes, and to problems involving both belt and drum rotating, like belt drives. It can be also used to systems
with the belt replaced by a rope or a string.
In case of V-shaped belts Eq. (4.8) should be replaced by
(4.9
)

where is the angle of the V of the belt cross section.


Fig. 4.8a shows an example of a hoisting machine lifting a load by means of a cable passing over an undriven sheave and wrapped
around a driven drum. The rotary motion of the drum drive is transformed into the translational motion of the cable by means of an
ideal transformer with the ratio 1 : r1, where r1 is the radius of the drum. The first and second part of the cable c1 and c2 is modeled by

elements characterizing the cable flexibility and damping. The element parameters of the second part of the cable c2 depend on the
length of c2. The friction between the cable and the sheave is represented by the belt-friction model given above. The sheave
dynamics, converted from rotary to translational motion along the cable, is represented by the elements of reduced parameters
(4.1
0)

where Js is the moment of inertia of the sheave, bs is the damping of its bearing, and r2 is its radius.

Figure 4.8 Hoisting machine.

Module 5
Mechanical planar systems
Module units

5.1

Planar motion of particles

5.1.1

Planar curvilinear motion

5.1.2

Planar motion in rectangular coordinates

5.1.3

5.1.2.1

Postulate of compatibility for planar motions

5.1.2.2

Postulate of continuity for planar forces

Multipole models

5.1.3.3

5.1.4

Example

Constrained planar motion of a particle

5.1.4.4

Constrained rectilinear motion

5.1.4.5

Example

5.1.4.6

Constrained curvilinear motion of a particle

5.1.4.7

Example

5.2

Planar motion of systems of particles

5.3

Systems of particles

5.3.1

Particles constrained by massless rod

5.3.1.1

Example

5.4

5.3.1.2

Example

5.3.1.3

Example

5.3.1.4

Example

5.3.2

Massless rod with polar variables

5.3.3

Planar motion in polar coordinates

5.3.4

Planar motion of coupled particles

5.3.5

Transformations

5.3.6

Planar motion of particles in path coordinates

Planar motion of multibody systems

5.4.1

Construction of free-body diagrams

5.5

Summary

5.6

Motion of bodies

5.6.1

Unconstrained and constrained motion

5.6.2

Systems of interconnected bodies

5.6.3

Analysis procedure

5.6.4

Translational, rotational and general plane motion

5.6.5

Mechanisms

5.6.6

Closed mechanisms

5.6.7

5.6.6.1

Vector loop equation for a four-bar linkage

5.6.6.2

Four-bar slider-crank equations

5.6.6.3

Inverted slider-crank equations

Systems of forces

Module overview.
We now begin our description of the motion of a particle along a curved path which lies in a single plane. Thus the vast majority of the
motions of points or particles encountered in engineering practice may be represented as plane motion. The case of plane motion
where all movement occurs in or can be represented as occurring in a single plane. A large proportion of the motions of machines and
structures in engineering can be represented as plane motion.
We begin our study of planar motion dynamic by first discussing in this module the motion of particles. A particle is a body whose
physical dimensions are so small compared with the radius of curvature of its path that we may treat the motion of the particle as that
of a point. For example, the wingspan of a jet transport flying between Los Angeles and New York is of no consequence compared with
the radius of curvature of its flight path, and thus the treatment of the airplane as a particle is an acceptable approximation.

The motion of particles (or rigid bodies) may be described by using coordinates measured from fixed reference axes (absolute motion
analysis) or by using coordinates measured from moving reference axes (relative-motion analysis). Both descriptions will be
developed and applied in the articles which follow.
If the particle is confined to a specified path, as with a bead sliding along a fixed wire, its motion is said to be constrained. If there are
no physical guides the motion is said to be unconstrained.
Module objectives. When you have completed this module you should be able to:
1. Understand the fundamental laws governing dynamics of mechanical planar systems.
2. Set up multipole diagrams representing dynamic models of these systems from pure physical elements and transducers.
3. Use the computer package DYNAST to submit multipole diagrams representing the mechanical planar systems in their
graphical form.
4. Use DYNAST to specify excitation and initial conditions of these models, invoke numerical time-domain analysis, and plot the
results.
Module prerequisites. Modules on dynamics of mechanical rectilinear and rotational systems of this course.

5.1 Planar motion of particles

5.1.1 Planar curvilinear motion


5.1.2 Planar motion in rectangular coordinates
5.1.3 Multipole models

5.1.4 Constrained planar motion of a particle

5.1.1 Planar curvilinear motion

A particle is a body whose physical dimensions are so small compared with the radius of curvature of its path that the motion of the
particle can be approximated as that of a point. A particle is in a planar motion if its path lies in a single plane. An example of a planar
path of a particle is given in Fig. 5.1a. Point A represents the particle position at time t. The particle position can be completely
specified by the position vector r measured from the fixed origin O. At time t + t, the particle is at point A, the position of which is
specified by the vector r + r.

Figure 5.1 (a) Particle position, (b) velocity, and (c) acceleration vectors.
The particle displacement during time t is indicated by the vector r representing the vector change of position. If an origin O is
chosen at some location different from O, the position vector r changes, but r remains unchanged. The distance travelled by the
particle as it moves along the path from A to A is the scalar length s measured along the curved path. Thus, the vector displacement
r is distinguished from the scalar distance s.

The instantaneous velocity v of the particle is defined as the limiting value of the particle average velocity as the time interval
approaches zero. Thus,

The direction of r approaches that of the tangent to the path as t approaches zero and, hence, the velocity v is always a vector
tangent to the path as shown in Fig. 5.1b.
The definition of the derivative of a scalar quantity can be extended to include a vector quantity so that

Thus the derivative of a vector is itself a vector having both a magnitude and a direction. The magnitude of v is called the speed and is
the scalar

In Fig. 5.1b, the velocity of the particle at A is denoted by the tangent vector v and the velocity at A by the tangent vector . There is
thus a vector change in the velocity during the time t. The velocity v at A plus the change v must equal the velocity at A. So we
may write
. Inspection of the vector diagram in Fig. 5.1c shows that v depends both on the change in magnitude of v and
on the change in direction of v.
The instantaneous acceleration a of the particle is defined as the limiting value of the average acceleration as the time interval
approaches zero, or

By definition of the derivative, then, we write

As the interval t approaches zero, the direction of the change v approaches that of the differential change dv and, hence, a.
Therefore, the acceleration a includes the effect of change in magnitude of v as well as the change of direction of v. It is apparent from
Fig. 5.1c that, in general, the direction of the acceleration of a particle in curvilinear motion is neither tangent nor normal to the particle
path.

5.1.2 Planar motion in rectangular coordinates

Fig. 5.2a shows again the plane in which the curvilinear path of a particle takes place. In addition, the rectangular coordinate axes x
and y are drawn on the plane. Such a coordinate system is often called Cartesian. The particle position represented by point A can now
be described at any time t by specifying the point coordinates xA and yA measured from the fixed origin of the axes O.

Figure 5.2 Rectangular components of particle (a) velocity and (b) acceleration.
The particle velocity v is decomposed in Fig. 5.2a into the rectangular components
particle and its x- and y-components

and

are shown in Fig. 5.2b. It is obvious that

. The acceleration a of the

as well as that

5.1.2.1 Postulate of compatibility for planar motions


Due to geometric continuity of space postulate of compatibility holds also for planar motions. Fig. 5.3a shows a particle displaced from
the origin O to point A, then to point B, and finally back to the origin O. Regardless of the shape of the actual closed path of the
particle through the points the total path length is zero according postulate of compatibility.
Regardless of the shape of the actual closed path of the particle through the points the total path length is zero according postulate of
compatibility.
Thus, in the vector form postulate
(5.1
)

Figure 5.3 Postulate of (a) compatibility and (b) continuity.


These relations represent geometric constraints which are placed on the planar motions of points
Postulate of compatibility

5.1.2.2 Postulate of continuity for planar forces


If a particle of mass m is isolated in the primary inertial system and is subjected to the action of a single force F then Newtons second
law of motion becomes a vector relation and may be written

where the acceleration a of the particle is always in the direction of the applied force.
This property is the inertia of the particle which is its resistance to rate of change of velocity.
Equation 3/3, or any one of the component forms of the force-mass-acceleration equation, is usually referred to as the equation of
motion.
(5.2
)
(5.3
)
(5.4
)

The vector sum


of () means the vector sum of all forces acting on the particle in question. Similarly, the corresponding scalar force
summation in any one of the component directions means the sum of the components of all forces acting on the particle in that
particular direction. The reliable way to account accurately and consistently for every force is to form the free-body diagram of the
particle under consideration isolating it from all contacting and influencing bodies and replacing the bodies removed by the forces they
exert on the particle isolated.

5.1.3 Multipole models

Figure 5.4 (a) Bob submodel, (b) bob symbol.


From the foregoing discussion it is obvious that the rectangular-coordinate representation of curvilinear motion is merely the
superposition of the components of two simultaneous rectilinear motions in the x- and y-directions. Therefore, everything covered in
the module on rectilinear motion may be applied separately to the x-motion and to the y-motion. Then the resulting curvilinear motion
is obtained by a vector combination of the x- and y-components of the position, velocity, acceleration and force vector.
Obviously, also the multipole model of a particle in a curvilinear motion can be decomposed into two rectilinear models, one for
dynamic interactions in the x direction, and the other one in the y direction. In most practical cases these two models will be coupled,
however.

5.1.3.3 Example
Fig. 5.5a shows the trajectory of a cannon ball represented by point A. The ball was launched from the position A0 at the elevation
angle with the initial velocity v0. Let us neglect for simplicity aerodynamic drag, the earths curvature and rotation of the ball. Then

the only external force acting on the ball is the vertical gravitational force Fg = mg, where m is the mass of the ball, and g is the
acceleration due to gravity.

Figure 5.5 Cannon ball (a) trajectory, and (b) multipole diagram.
In the two-part multipole diagram given in Fig. 5.5c, nodes Ax and Ay represent the ball motion in the x- and y-direction, respectively.
Each of the nodes is associated with the inertor modeling the ball inertia effect in the related direction. The gravity effect is respected
by the source of force Fg acting at node Ay. Integrators are added to the diagram to compute the ball position in the x- and y-direction.
To compute, for example, the magnitude of the tangential velocity vA of the ball, or the distance sA travelled by the ball along its
trajectory, we could add to the model the following expressions

either in a textual or block-diagram form.


An important application of two-dimensional dynamic theory is the problem of projectile motion. For a first treatment of the subject,
we neglect aerodynamic drag and the earths curvature and rotation, and we assume that the altitude range is small enough so that the

acceleration due to gravity can be considered constant. With these assumptions, rectangular coordinates are useful for the trajectory
analysis. For the axes shown in Fig. 5.5, the acceleration components are

Integration of these accelerations yields

In all these expressions, the subscript zero denotes initial conditions, frequently taken as those at launch where, for the case illustrated,
x0 = y0 = 0. Note that the quantity g is taken to be positive throughout this text.
We can see that the x- and y-motions are independent for the simple projectile conditions under consideration. Elimination of the time t
between the x- and y-displacement equations shows the path to be parabolic (see Sample Problem 2/6). If we were to introduce a drag
force which depends on the speed squared (for example), then the x- and y-motions would be coupled (interdependent), and the
trajectory would be nonparabolic.
When extreme accuracy is to be achieved in describing projectile motion when large velocities and high altitudes are involved, the
shape of the projectile, the variation of g with altitude, the variation of the air density with altitude, and the rotation of the earth must
all be accounted for.

5.1.4 Constrained planar motion of a particle

Until now, we have considered unconstrained motion where the particle was free of any mechanical guides and followed a path
determined by its initial motion and by the forces which were applied to it from external sources. Now we have come to discussing
constrained motion where the path of the particle is partially or totally determined by restraining guides.

5.1.4.4 Constrained rectilinear motion


Let us first consider a guide constraining the particle motion to the rectilinear trajectory shown in Fig. 5.6a. The particle is represented
by point A. Such a constraint is imposed, for example, on the motion of a train following a straight segment of the railroad.

Figure 5.6 Particle (a) rectilinear trajectory, (a)


The particle rectilinear motion along the trajectory shown in Fig. 5.6a can be described in rectangular coordinates by the relation
(5.5
)

By differentiating (5.5) with respect to time we obtain the relationship between the x and y components of the particle velocity vA along
the straight trajectory
(5.6
)

Let us assume that the particle can move along the constrained trajectory freely without any friction. Then the only force imposed on
the particle by the constraining guide is the force Fn. It is normal to the particle trajectory as it is illustrated in Fig. 5.6b.
As the external forces exerted on the particle are not known yet, neither the velocity vA, nor the force Fn can be evaluated so far.
However, it is obvious from the comparison of Fig. 5.6a and b that the magnitudes of the FAx and FAy components of Fn are interrelated
as
(5.7
)

Rewriting (5.6) and (5.7) in a common matrix notation


(5.8
)

Reveals that this is the constitutive relation of a pure transformer with the ratio n = 1/k. Recollect that a pure transformer neither
dissipates nor accumulates any energy. To verify this you may substitute for vAx and FAy from (5.8) to the equation for the total energy
inflow into the transformer
(5.9
)

5.1.4.5 Example
Fig. beadstra shows the free-body diagram of a bead sliding along a fixed straight wire. There are two external forces acting on the
bead: the gravitational force Fg and the reaction force Fn due to the wire constraint. To draw the diagram we must start with the vertical
vector Fg. Its magnitude is mg, where m is the particle mass, and g is the gravitational acceleration. In the next step, we are able to
draw the parallelogram determining magnitudes of the forces Fn and FA. Note, that the x and y components of FA can be determined
using (5.7).

Figure 5.7
Instead of drawing the parallelogram we can directly construct the corresponding multipole model shown in Fig. 5.7b. The transformer
ratio n = 1/k is of a negative value respecting the down-going slope of the constraint.

5.1.4.6 Constrained curvilinear motion of a particle


A guide constraining the motion of a particle represented by point A to the curvilinear trajectory can be described by the equation
(5.1
0)

where f(.) is a time-invariant function. After differentiating (5.10) we receive the linearized relation between the x and y components of
the instantaneous velocity vA
(5.1
1)

Following then the same reasoning procedure which we used for the rectilinear constraint we come to the conclusion that also the
curvilinear constraint can be modelled by a pure transformer. Its ratio is
(5.1
2)

Note that the ratio changes with the particle position.

5.1.4.7 Example
A curvilinear constraint is characterized by the relation

Therefore, the ratio of the transformer representing the constraint is

5.3 Systems of particles

5.3.1 Particles constrained by massless rod


5.3.2 Massless rod with polar variables
5.3.3 Planar motion in polar coordinates
5.3.4 Planar motion of coupled particles
5.3.5 Transformations
5.3.6 Planar motion of particles in path coordinates

5.3.1 Particles constrained by massless rod

Fig. 5.8a shows two particles moving in the x-y plane coupled by a massless rigid rod. The particles are represented by points A and B.
Let us determine the constrains imposed by the rod on the particle motion.

Figure 5.8 (a) Massless rod, (b) its model.

The geometric configuration of the rod is characterized by the relation


(5.1
3)

where is the rod length and

are projections of the rod onto the x and y axes. After differentiating (5.13) with respect to time
(5.1
4)

The rod must be in equilibrium, i.e. all forces and moments applied to it must be in balance. The balance of forces can be written in the
form
(5.1
5)

Equilibrium of moments can be expressed as


(5.1
6)

After substituting for FAx and FAy from (5.15) balance of moments becomes

(5.1
7)

If we rewrite (5.14) and (5.17) in a common matrix form


(5.1
8)

we can conclude that the constraint provided by the massless rod can be modelled by a pure transformer with the ratio n = yBA/xBA
shown in Fig. 5.8b.

5.3.1.1 Example

Figure 5.9 (a) Mathematical pendulum, (b) its dynamic model.

5.3.1.2 Example

Figure 5.10 (a) Mathematical double pendulum, (b) its dynamic model.

5.3.1.3 Example

Figure 5.11 (a) Carriage with pendulum, (b) its dynamic model.

5.3.1.4 Example

Figure 5.12 (a) Sliding rod, (b) its dynamic model.

5.3.2 Massless rod with polar variables

Figure 5.13
(5.1
9)
(5.2
0)

M = Mx + My

(5.2
1)
(5.2
2)

5.3.3 Planar motion in polar coordinates


(5.2
3)

where r is the length of vector r and is the angle between this vector and the axis x.
By differentiating these relation with respect to time we will obtain velocities of point A in the coordinate system x, y expressed in
terms of velocities in the coordinate system ,
(5.24)
(5.25)
(5.26)
(5.27)
(5.2
8)

where

(5.2
9)

and
(5.3
0)

5.3.4 Planar motion of coupled particles

Table 5.1 Massless rods


Coupling
Translatorytorotary

Mechanism

Pure model

Translatorytorotary

5.3.5 Transformations

An example of the magnitude of the error introduced by neglect of the motion of the earth may be cited for the case of a particle that is
allowed to fall from rest (relative to the earth) at a height h above the ground. We can show that the rotation of the earth gives rise to an
eastward acceleration (Coriolis acceleration) relative to the earth and, neglecting air resistance, that the particle falls to the ground a
distance

east of the point on the ground directly under that from which it was dropped. The angular velocity of the earth is = 0.729(104)
rad/s, and the latitude, north or south, is . At a latitude of
and from a height of 200 m, this eastward deflection would be x = 43.9
mm.
These corrections are negligible for most engineering problems which involve the motions of structures and machines on the surface of
the earth. In such cases, the accelerations measured with respect to reference axes attached to the surface of the earth may be treated as
"absolute," and Eq. 3/2 may be applied with negligible error to experimental measurements made on the surface of the earth.* If the
ideal experiment described were performed on the surface of the earth and all measurements were made relative to a reference system
attached to the earth, the measured results would show a slight discrepancy upon substitution into Eq. 3/2. This discrepancy would be

due to the fact that the measured acceleration would not be the correct absolute acceleration. The disc The discrepancy would
disappear when we introduced the corrections due to the acceleration components of the earth.

5.3.6 Planar motion of particles in path coordinates

The motion of A may also be described by measurements along the tangent t and normal n to the curve. The direction of n lies in the
local plane of the curve. These last measurements are known as path variables.

5.4 Planar motion of multibody systems

When a body is in equilibrium, the resultant of all forces acting on it is zero. Thus, the resultant force R and the resultant couple M are
both zero, and we have the equilibrium equations
R = SF = O fR = 2F = O M = 2M = O )
These requirements are both necessary and sufficient conditions for equilibrium.
Once we reach a decision about which body or combination of bodies is to be analyzed, then this body or combination treated as a
single body is isolated from all surrounding bodies. This isolation is accomplished by means of the free-body diagram, which is a
diagrammatic representation of the isolated body or combination of bodies treated as a single body, showing all forces applied to it by
mechanical contact with other bodies which are imagined to be removed. If appreciable body forces are present, such as gravitational
or magnetic attraction, then these forces must also be shown on the diagram of the isolated body.
We noted that forces are applied both by direct physical contact and by remote action and that forces may be either internal or external
to the body under consideration. We further observed that the application of external forces is accompanied by reactive forces and that

both applied and reactive forces may be either concentrated or distributed. Additionally the principle of transmissibility was
introduced, which permits the treatment of force as a sliding vector as far as its external effects on a rigid body are concerned.
Figure 3/1 shows the common types of force application on mechanical systems for analysis in two dimensions. In each example the
force exerted on the body to be isolated by the body to be removed is indicated. Newtons third law, which notes the existence of an
equal and opposite reaction to every action, must be carefully observed. The force exerted on the body in question by a contacting or
supporting member is always in the sense to oppose the movement of the body which would occur if the contacting or supporting
member were removed.

5.4.1 Construction of free-body diagrams


The following steps are involved.
1. A clear decision is made concerning which body or combination of bodies is to be isolated. The body chosen will usually
involve one or more of the desired unknown quantities.
2. The body or combination chosen is next isolated by a diagram which represents its complete external boundary. This boundary
defines the isolation of the body from all other contacting or attracting bodies, which are considered removed. This step is often
the most crucial of all. We should always be certain that we have completely isolated the body before proceeding with the next
step.
3. All forces which act on the isolated body as applied by the removed contacting and attracting bodies are next represented in
their proper positions on the diagram of the isolated body. A systematic traverse of the entire boundary will disclose all contact
forces. Weights, where appreciable, must be included. Known forces should be represented by vector arrows, each with its
proper magnitude, direction, and sense indicated. Each unknown force should be represented by a vector arrow with the
unknown magnitude or direction indicated by symbol. If the sense of the vector is also unknown, it may be arbitrarily assumed.
The calculations will reveal a positive quantity if the correct sense was assumed and a negative quantity if the incorrect sense

was assumed. It is necessary to be consistent with the assigned characteristics of unknown forces throughout all of the
calculations.
4. The choice of coordinate axes should be indicated directly on the diagram.

5.5 Summary
To set up a dynamic model of a real rectilinear system and to simulate its behavior using DYNAST, you should proceed through the
following steps.
1. Sketch the geometric configuration of your system, i.e.:
o clearly separate the system of your concern from the system surroundings and determine a reference frame for the
system
o associate the rectilinear motions with a coordinate axis
o for each body denote the geometric point representing its rectilinear motion and give it a name
o assume the positive velocity of each such point in the direction of the chosen coordinate axis
2. Set up a multipole diagram representing the dynamic model of your system and its surroundings (you may use directly the
DYNAST schematic editor):
o for each body place a node into the diagram corresponding to the geometric point representing the body velocity
o give each node a name identical with the name of the point (using a label in the DYNAST schematic editor)
o associate each of the nodes with the symbol for inertors

o place into the diagram one or more symbols for reference nodes representing the reference frame
o place symbols for elements representing couplings between the nodes (if necessary, interconnect the element pins with
the nodes by links)
o add symbols for elements representing the system surroundings
o invert the element symbols if their velocity is oriented in the opposite direction with respect to the chosen coordinate
axis
o give names to the inertors and the other elements and specify values of their parameters
3. Specify analysis of your system model:
o select the Numerical Nonlinear Analysis
o specify the required Time Interval of the analysis and set the values of Initial Conditions
o select Desired Variables like node and element velocities, element forces, or block output nodes
o To invoke the analysis, select Run Analysis & Plot in menu Run
4. Interpret the analyzed results:
o Set the desired variables in menu Plot
o in menu Axes choose the way you want to plot the variables
o the assumed positive polarity of velocities are oriented in the direction of the chosen coordinate axes whereas the
assumed positive polarity of forces of the elements is opposite to this direction

5.6 Motion of bodies

In Chapter on particle dynamics, we developed the relationships governing the displacement, velocity, and acceleration of points as
they moved along curved paths. In rigid-body dynamics we use these same relationships but must also account for the rotational
motion of the body as well. Thus, rigid-body dynamics involves both linear and angular quantities.
A rigid body was defined in the previous chapter as a system of particles for which the distances between the particles remain
unchanged. This formulation is, of course, an ideal one since all solid materials change shape to some extent when forces are applied to
them. Nevertheless, if the movements associated with the changes in shape are very small compared with the overall movements of the
body as a whole, then the ideal concept of rigidity is quite acceptable.
A rigid body executes plane motion when all parts of the body move in parallel plane; for convenience, we generally consider the plane
of motion to be the plane that contains the mass center. This idealization fits a very large category of rigid-body motions encountered
in engineering. The plane motion of a rigid body may be divided into several categories, as represented m Fig. 5.14.

Figure 5.14 Rigid-body plane motions.


A body which has appreciable dimensions normal to the plane of motion but is symmetrical about that plane of motion through the
mass center may be treated as having plane motion. This idealization clearly fits a very large category of rigid-body motions.
Basic to the approach to kinetics is the isolation of the body or system to be analyzed. For problems involving the instantaneous
relationships among force, mass, and acceleration or momentum, the body or system should be explicitly defined by isolating it with
its free-body diagram.
In the kinetics of rigid bodies which have angular motion, it is necessary to introduce a property of the body which accounts for the
radial distribution of its mass with respect to a particular axis of rotation normal to the plane of motion. This property is known as the
mass moment of inertia of the body, and it is essential that we will be able to calculate this property in order to solve rotational
problems.
The force equation,
(5.3
1)

tells us that the resultant


of the external forces acting on the body equals the mass m of the body times the acceleration a of its
mass center G. The moment equation taken about the mass center,
(5.3
2)

tells us that the resultant moment about the mass center of the external forces on the body equals the time rate of change of the angular
momentum of the body about the mass center.

We recall from our study of statics that a general system of forces acting on a rigid body may be replaced by a resultant force applied at
a chosen point and a corresponding couple. By replacing the external forces by their equivalent force-couple system with the resultant
force acting through the mass center, we may then visualize the action of the forces and the corresponding dynamic response of the
body with the aid of Fig. 5.15. The a-part of the figure represents the free-body diagram of the body. The b-part of the figure shows the
equivalent force-couple system with the resultant force applied through G. The c-part of the figure represents the resulting dynamic
effects as specified by Eqs. 4/1 and 4/9 and is termed here the kinetic diagram. The equivalence between the free body diagram and the
kinetic diagram enables us to clearly visualize and easily remember the separate translational and rotational effects of the forces
applied to a rigid body.

Figure 5.15 Equivalent force-couple system.


We now apply the foregoing relationships to the case of plane motion. Fig. 5.16 represents a rigid body moving with plane motion in
the x-y plane. The mass center

Figure 5.16 Equivalent force-couple system.


The results embodied in our basic equations of motion for a rigid body in plane motion, Eqs. 6/1, are represented diagrammatically in
Fig. 6/4, which is the two-dimensional counterpart of the a- and c-parts of Fig. 6/1 for a general three-dimensional body. The free-body
diagram discloses the forces and moments appearing on the left-hand side of our equations of motion. The kinetic diagram discloses
the resulting dynamic response in terms of the translational term ma and the rotational term la which appear on the right-hand side of
Eqs. 6/1. As previously mentioned, the translational term ma will be expressed by its x-y, n-t, or r-6 components once the appropriate

inertial reference system is designated. The equivalence depicted in Fig. 6/4 is basic to our understanding of the kinetics of plane
motion and will be employed frequently in the solution of problems.

Figure 5.17 6/4

Figure 5.18 ...

5.6.1 Unconstrained and constrained motion

The motion of a rigid body may be unconstrained or constrained. The rocket moving in a vertical plane, Fig. 6/6a, is an example of
unconstrained motion as there are no physical confinements to its motion. The two components ax and ay of the mass-center
acceleration and the angular acceleration a may be determined independently of one another by direct application of Eqs. 6/1. The bar
in Fig. 6/66, on the other hand, represents a constrained motion, where the vertical and hori2ontal guides for the ends of the bar impose

a kinematic relationship between the acceleration components of the mass center and the angular acceleration of the bar. Thus, it is
necessary to determine this kinematic relationship established in ... .

Figure 5.19 6/6


In general, dynamics problems which involve physical constraints to motion require a kinematic analysis relating linear to angular
acceleration before the force and moment equations of motion can be solved.

5.6.2 Systems of interconnected bodies

Upon occasion, in problems dealing with two or more connected rigid bodies whose motions are related kinematically, it is convenient
to analyze the bodies as an entire system. Figure 6/7 illustrates two rigid bodies hinged at A and subjected to the external forces shown.
The forces in the connection at A are internal to the system and are not disclosed.
The resultant of all external forces must equal the vector sum of the two resultants m1a1 and m2a2, and the sum of the moments about
some arbitrary point such as P of all external forces must equal the moment of the resultants, I11 + I22 + m1a1d1 + m2a2d2. Thus, we
may state

where the summations on the right-hand side of the equations represent as many terms as there are separate bodies.
If there are more than three remaining unknowns in a system, however, the three independent scalar equations of motion, when applied
to the system, are not sufficient to solve the problem. In this case, more advanced methods such as virtual work (Art. 6/7) or
Lagranges equations (not discussed in this book*) could be employed or else the system dismembered and each part analyzed separate
resulting equations solved simultaneously.

Figure 5.20 6/7

Figure 5.21 ...

Figure 5.22 ...

Figure 5.23 ...

Figure 5.24 ...

Figure 5.25 ...

5.6.3 Analysis procedure

In the solution of force-mass-acceleration problems for the plane motion of rigid bodies, the following steps should be taken after the
conditions and requirements of the problem are clearly in mind.

First, identify the class of motion and then solve for any needed linear and angular accelerations which can be determined solely from
given kinematic information. In the case of constrained plane motion, it is usually necessary to establish the relation between the linear
acceleration of the mass center and the angular acceleration of the body by first solving the appropriate relative-velocity and relativeacceleration equations.
Always draw the complete free-body diagram of the body to be analyzed. Assign a convenient inertial coordinate system and label all
known and unknown quantities.
Apply the three equations of motion from Eqs. 6/1, being consistent with the algebraic signs in relation to the choice of reference axes.
Equation 6/2 or 6/3 may be employed as an alternative to one of Eqs. 6/1. Combine these relations with the results from any needed
kinematic analysis. Count the number of unknowns and be certain that there are an equal number of independent equations available.
For a solvable rigid body problem in plane motion, there can be no more than the five scalar unknowns which can be determined from
the three scalar equations of motion, obtained from Eqs. 6/1, and the two scalar component relations that come from the relativeacceleration equation.

5.6.4 Translational, rotational and general plane motion

The foregoing developments will now be applied to the three cases of motion in a plane, namely, translation, fixed-axis rotation, and
general plane motion in the three articles which follow.
Rigid-body translation in plane motion was described in Art. 5/1 and illustrated in Figs. 5/la and 5/lb, where we saw that every line in a
translating body remains parallel to its original position at all times. In rectilinear translation all points move in straight lines, whereas
in curvilinear translation all points move on congruent curved paths. In either case, there is no angular motion of the translating body,
so that both CD and a are zero. Therefore, from the moment relation of Eqs. 6/1, we see that all reference to the moment of inertia is
eliminated for a translating body.
For a translating body, then, our general equations for plane motion, Eqs. 6/1, may be written

For rectilinear translation, illustrated in Fig. 6/8a, if the x-axis is chosen in the direction of the acceleration, then the two scalar force
equations become

and

For curvilinear translation, Fig. 6/8b, if we use n-t coordinates, the two scalar force equations become
both cases,
.

and

. In

Figure 5.26 ...


Rotation of a rigid body about a fixed axis O was described in Art. 5/2 and illustrated in Fig. 5/lc. For this motion, we saw that all
points in the body describe circles about the rotation axis, and all lines of the body in the plane of motion have the same angular
velocity w and angular acceleration a. The acceleration components of the mass center for circular motion are most easily expressed in
n-t coordinates, so we have an = r2 and at = r, as shown in Fig. 6/9a for rotation of the rigid body about the fixed axis through O.
The b-part of the figure represents the free-body diagram, and the equivalent kinetic diagram in the c-part of the figure shows the force
resultant ma in terms of its n- and t-components and the resultant couple I. Our general equations for plane motion, Eqs. 6/1, are
directly applicable and are repeated here.
Thus, the two scalar components of the force equation become
and
. In applying the moment equation about
G, it is essential to account for the moment of the force applied to the body at O, so this force must not be omitted from the free-body
diagram.
For fixed-axis rotation, it is generally useful to apply a moment equation directly about the rotation axis O. We derived this equation
previously as Eq. 6/4, which is repeated here.
(5.3
5)

From the kinetic diagram in Fig. 6/9c, we may obtain Eq. 6/4 very easily by evaluating the moment of the resultants about O, which
becomes
. Substitution of the transfer-of-axis relation for mass moments of inertia, IO = I + mr2, gives
.
For the common case of rotation of a rigid body about a fixed axis through its mass center G, clearly, a = 0, and therefore
resultant of the applied forces then is the couple I.

. The

We may combine the resultant force mat and resultant couple I by moving mat to a parallel position through point Q on line OG, Fig.
6/10, located by mrq = I + mr(r). Using the transfer-of axis theorem and
gives
. Point Q is called the center of
percussion and has the unique property that the resultant of all forces applied to the body must pass through it. It follows that the sum
of the moments of all forces about the center of percussion is always zero,

Figure 5.27 ...

Figure 5.28 ...


The dynamics of general plane motion of a rigid body combines translation and rotation. In Art. 6/2 we represented such a body in Fig.
6/4 with its free-body diagram and its kinetic diagram, which discloses the dynamic resultants of the applied forces. Figure 6/4 and
Eqs. 6/1, which apply to general plane motion, are repeated here for convenient reference.
As previously, the two scalar components of the force equation are expressed in whatever coordinate system most readily describes the
acceleration of the mass center. Direct application of these equations expresses the equivalence between the externally applied forces,
as disclosed by the free-body diagram, and their force and moment resultants, as represented by the kinetic diagram.
In Art. 6/2 we also showed, with the aid ofFig. 6/5, the application of the alternative relation for moments about any point Vi Eq. 6/2.
This figure and this equation are also repeated here for easy reference.
(5.3
6)

In some instances, it may be more convenient to use the alternative moment relation of Eq. 6/3 when moments are taken about a point
P whose acceleration is known. It is also noted that the equation for moments about a nonaccelerating point O on the body, Eq. 6/4,
constitutes still another alternative moment relation and at times may be used to advantage.
In working a problem in general plane motion, we first observe whether the motion is unconstrained or constrained, as illustrated in the
examples of Fig. 6/6. If the motion is constrained, we must account for the kinematic relationship between the linear and the angular
accelerations and incorporate it into our force and moment equations of motion. If the motion is unconstrained, the accelerations can
be determined independently of one another by direct application of the three motion equations, Eqs. 6/1.
In order for a rigid-body problem to be solvable, the number of unknowns cannot exceed the number of independent equations
available to describe them, and a check on the sufficiency of the relationships should always be made. At the most, for plane motion
we have three scalar equations of motion and two scalar components of the vector relative-acceleration equation for constrained
motion. Thus, we can handle as many as five unknowns.
In formulating the solution to a problem, we recognize that the directions of certain forces or accelerations may not be known at the
outset, so that it may be necessary to make initial assumptions whose validity will be proved or disproved when the solution is carried
out. It is essential, however, that all assumptions made be consistent with the principle of action and reaction and with any kinematic
requirements, which are also called conditions of constraint.

Figure 5.29 ...

5.6.5 Mechanisms
Kinematics The study of motion without regard to forces.
Kinetics The study of forces on systems in motion.
These two concepts are really not physically separable. We arbitrarily separate them for instructional reasons in engineering education.
It is also valid in engineering design practice to first consider the desired kinematic motions and their consequences, and then
subsequently investigate the kinetic forces associated with those motions. The student should realize that the division between
kinematics and kinetics is quite arbitrary and is done largely for convenience. One cannot design most dynamic mechanical systems
without taking both topics into thorough consideration.

A mechanism is a device which transforms motion to some desirable pattern and typically develops very low forces and transmits little
power. A machine typically contains mechanisms which are designed to provide significant forces and transmit significant power.
Some examples of common mechanisms are a pencil sharpener. a camera shutter, an analog clock, a folding chair, an adjustable desk
lamp, and an umbrella. Some examples of machines which posses motions similar to the mechanisms listed above are a food blender, a
bank vault door, an automobile transmission, a bulldozer, a robot, and an amusement park ride.
Any mechanical system can be classified according to the number of degrees of freedom (DOF) which it possesses. The systems DOF
is equal to the number of independent parameters measurements) which are needed to uniquely define its position in space at any
instant of time.
Degree of freedom of a system is defined as the number of independent coordinates required to define its position.
Linkages are the basic building blocks of all mechanisms. We will show in later chapters that all common forms of mechanisms (cams,
gears, belts, chains) are in fact variations of linkages. Linkages are made up from links and joints.
A link, as shown in Figure 2-2, is an (assumed) rigid body which possesses at least two nodes which are points for attachment with
other links.
Binary link one with two nodes.
Ternary link one with three nodes.
Quaternary link one with four nodes.
A joint is a connection between two or more links (at their nodes, which allows some motion, or potential motion, between the
connected links. Joints (also called kinematic pairs) can be classified in several ways:
1. By the type of contact between the elements, line point or surface.
2. By the number of degrees of freedom allowed at the point.

3. By the type of physical closure of the joint: either force or form closed.
4. By the number of links joined (order of the joint).
A kinematic chain is defined as an assemblage of links and joints, interconnected in a way to provide a controlled output motion in
response to a supplied input motion.
A mechanism is defined as a kinematic chain in which at least one link has been "grounded," or attached, to the frame of reference
(which itself may be in motion).
A machine is defined as a combination of resistant bodies arranged to compel the mechanical forces of nature to do work accompanied
by determinate motions.
A machine is a collection of mechanisms arranged to transmit forces and do work.
We will now define a crank as a link which makes a complete revolution and is pivoted to ground, a rocker as a link which has
oscillatory (back and forth) rotation and is pivoted to ground, and a coupler (or connecting rod) which has complex motion and is not
pivoted to ground. Ground is defined as any link or links that are fixed (non-moving) with respect to the reference frame. Note that the
reference frame may in fact itself be in motion.
Kinematic chains or mechanisms may be either open or closed. Figure 2-4 shows both open and closed mechanisms. A common
example of an open mechanism is an industrial robot.
Pins in a body of unit length are constrained to slide in horizontal and vertical slots as shown in Fig. P2.4

5.6.6 Closed mechanisms

5.6.6.1 Vector loop equation for a four-bar linkage


The links of the four-bar mechanism in Figure 4-6 are replaced by position vectors. The directions of the position vectors are chosen so
as to define their angles where we desire them to be measured. By definition, the angle of a vector is always measured at its root, not
not at its head.
These choices of vector directions and senses, as indicated by their arrowheads, lead to this vector loop equation:
(5.3
7)

This notation for the position vectors uses the labels of the points at the vector tips and roots (in that order) as subscripts. The second
subscript is conventionally omitted if it is the origin of the global coordinate system (point O).
This equation can now be separated into its x and y parts and each set to zero.
x component:
(5.3
8)

but: 1 = 0, so:
(5.3
9)

y component:

(5.4
0)

but: 1 = 0, so:
(5.4
1)

5.6.6.2 Four-bar slider-crank equations


The same vector loop approach as used above can be applied to a linkage containing sliders. Figure 4-9 shows an offset fourbar slidercrank linkage, inversion 1. The term offset means that the slider axis extended does not pass through the crank pivot. This is the
general case. (The nonoffset slider-crank linkages shown in Figure 2-13 (p. 44) are the special cases.) This linkage could be
represented by only three position vectors,
will be easier to use four vectors,
pair of vectors

and

,
, and

, and
with

, but one of them (

) will be a vector of varying magnitude and angle. It

arranged parallel to the axis of sliding and

are orthogonal components of the position vector

perpendicular. In effect the

from the origin to the slider.

This particular arrangement of position vectors leads to a vector loop equation similar to the pin-jointed fourbar example:
(5.4
2)

Compare equation 4.14a to equation 4.5a (p. 156) and note that the only difference is the sign of
Separate the real and imaginary components:
real part (x component):

(5.4
3)

but: 1 = 0, so:
(5.4
4)

imaginary part (y component):


(5.4
5)

but: 1 = 0, so:
(5.4
6)

5.6.6.3 Inverted slider-crank equations


Figure 4-lOa shows inversion 3 of the common fourbar slider-crank linkage in which the sliding joint is between links 3 and 4 at point
B. This is shown as an offset slider-crank mechanism. The slider block has pure rotation with its center offset from the slide axis.
(Figure 2-13c, p. 44, shows the nonoffset version of this linkage in which the vector

is zero.)

The global coordinate system is again taken with its origin at input crank pivot O2 and the positive x axis along link 1, the ground link.
A local axis system has been placed at point B in order to define . Note that there is a fixed angle within link 4 which defines the
slot angle with respect to that link.

In Figure 4-lOb the links have been represented as position vectors having senses consistent with the coordinate systems that were
chosen for convenience in defining the link angles. This particular arrangement of position vectors leads to the same vector loop
equation as the previous slider-crank example. Equations 4.14 and 4.1 apply to this inversion as well. Note that the absolute position of
point B is defined by vector
vector difference

which varies in both magnitude and direction as the linkage moves. We choose to represent

as the

in order to use the actual links as the position vectors in the loop equation.

All slider linkages will have at least one link whose effective length between joints will vary as the linkage moves. In this example the
length of link 3 between points A and B, designated as b, will change as it passes through the slider block on link 4. Thus the value of
b will be one of the variables to be solved for in this equation. Another variable will be , the angle of link 4. Note however, that we
have an unknown in , the angle of link 3. This is a total of three unknowns. Equation 4.l5 can only be solved for two unknowns. Thus
we require another equation to solve the system. There is a fixed relationship between angles and , shown as in Figure 4-10,
which gives the equation:
(5.4
7)

where the + sign is used for the open and the sign for the crossed configuration.
Substituting equation 4.18 into equations 4.15 yields:

5.6.7 Systems of forces

A force has been defined as the action of one body on another. We find that force is a vector quantity, since its effect depends on the
direction as well as on the magnitude of the action and since forces may be combined according to the parallelogram law of vector
combination. Thus, the complete specification of the action of a force must include its magnitude, direction, and point of application,
in which case it is treated as a fixed vector.
The action of a force on a body can be separated into two effects, external and internal. Forces external to a body are then of two kinds,
applied forces and reactive forces. In dealing with the mechanics of rigid bodies, where concern is given only to the net external effects
of forces, experience shows us that it is not necessary to restrict the action of an applied force to a given point. This conclusion is
described by the principle of transmissibility, which states that a force may be applied at any point on its given line of action without
altering the resultant effects of the force external to the rigid body on which it acts. When only the resultant external effects of a force
are to be investigated, the force may be treated as a sliding vector, and it is necessary and sufficient to specify the magnitude, direction,
and line of action of the force.
The characteristic of a force expressed by Newtons third law must be carefully observed. The action of a force is always accompanied
by an equal and opposite reaction. It is essential for us to fix clearly in mind which force of the pair is being considered. The answer is
always clear when the body in question is isolated and the force exerted on that body (not by the body) is represented. It is very easy to
make a careless mistake and consider the wrong force of the pair unless we distinguish carefully between action and reaction.
Two forces F1 and F2 that are concurrent may be added by the parallelogram law in their common plane to obtain their sum or resultant
R as shown in Fig. 2/3a. If the two concurrent forces lie in the same plane but are applied at two different points as in Fig. 2/3b, by the
principle of transmissibility we may move them along their lines of action and complete their vector sum R at the point of concurrency
A. The resultant R may replace F1 and F2 without altering the external effects on the body upon which they act. Mathematically the
sum of the two forces may be written by the vector equation
(5.5
0)

In addition to the need for combining forces to obtain their resultant, we often have occasion to replace a force by its vector
components which act in specified directions. By definition, the two or more vector components of a given vector must vectorially add
to yield the given vector. Thus, the force R in Fig. 2/3a may be replaced by or resolved into two vector components F1 and F2 with the
specified directions merely by completing the parallelogram as shown to obtain the magnitudes of F1 and F2.
A special case of addition is presented when the two forces F1 and F2 are parallel, Fig. 2/4. They may be combined by first adding two
equal, opposite, and collinear forces F and F of convenient magnitude, which taken together produce no external effect on the body.
Adding F1 and F to produce R1 and combining with the sum R2 of F2 and F yield the resultant R correct in magnitude, direction, and
line of action. This procedure is also useful in obtaining a graphical combination of two forces that are almost parallel and hence have
a point of concurrency which is remote and inconvenient.

Module 6
Formulation of system equations
Module units

6.1

Full set of equations

6.2

Reduced equations

6.2.1

Nodal formulation

6.2.1.1

Examples

6.2.2

Loop formulation

6.2.2.2

6.3

6.4

Example

Extended nodal formulation

6.3.1

Principles of the method

6.3.2

Stamp submatrices of uncontrolled elements

6.3.3

Stamp matrices of controlled physical elements

6.3.4

Stamp matrices of transducers

6.3.5

Stamp matrices of blocks

Block diagrams

6.5

6.4.1

Construction of block diagrams

6.4.2

Block diagrams vs. multipoles

Lagranges equations of multipoles

6.5.1

Example: Body with two revolute joints

6.5.2

Example: Flexible crank-and-slide mechanism

Module overview.
This module will introduce you into systematic procedures for formulation of equations for multidisciplinary linear dynamic systems
modelled by multipole or block-diagram models.
Module objectives. When you have completed this module you should be able to:
1. Understand the principles of equation formulation for dynamic systems.
2. Set up equations for a given system systematically using the method of stamp matrices.

3. Use the computer package DYNAST to submit the equations and to solve them.
Module prerequisites. The module of this course focused on multipole modeling in the domain of your interest.

6.1 Full set of equations

Values or waveforms of variables in a system multipole model can be obtained by solving a set of equations characterizing the model.
To formulate such equations, we must take into our consideration the following three types of relations.
Table 6.1 Examples of simple dynamic systems and their multipole models.
Configurations of real systems

Multipole diagrams

Part

Real parts

Physical elements

electrical energy supply

source of electrical current

electrical condenser

electrical capacitor

coil of wire

electrical inductor

electrical resistor

electrical conductor

Part

Real parts

Physical elements

water pump driven by motor

source of volume flow-rate

open water tank

fluid capacitor

long pipe

fluid inductor

water-flow restriction

fluid conductor

Part

Real parts

Physical elements

weight suspended on cable via pully

source of force

mass of moving body

inertor

helical spring

pure spring

oil-filled damper

pure damper

Part

Real parts

Physical elements

combustion engine

source of torque

flywheel

rotational inertor

long flexible shaft

torsional spring

ventilator propeller

rotational damper

(a) Constitutive relations characterizing each of the individual multipoles in the model.

(b) Postulate of continuity reflecting the physical laws governing the balance of energy and mass at the individual sites of interaction
in the system. The laws are given for different physical domains in Table 6.2. According to this postulate the sum of all the elemental
through variables ik at a node
(6.1
)

assuming that ik > 0 if ik leaves the node, and ik < 0 if ik enters the node.
(c) Postulate of compatibility resulting from the laws respecting the geometric connectedness of real systems given in Table 6.2.
According to this postulate the sum of all the elemental across variables vk along a closed oriented loop in the diagram
(6.2
)

where vk > 0 if the vk direction is in agreement with the loop orientation, and vk < 0 if the vk direction is opposite to the orientation of
the loop.
Table 6.2 Postulates of continuity and compatibility.
Physical domain

Postulate of continuity

Postulate of compatibility

electrical

general Kirchhoffs current law

Kirchhoffs voltage law

magnetic

continuity of magnetic flux

Amperes circuital law

preservation of energy

zeroth law of thermodynamics

thermal

fluid or acoustic
mechanical

conservation of mass

principle of pressure composition

dynamic equilibrium of forces

principle of motion composition

Table 6.3 Examples of simple dynamic systems and their multipole models.
Configurations of real systems

Multipole diagrams

Part

Real parts

Physical elements

electrical energy supply

source of electrical current

electrical condenser

electrical capacitor

coil of wire

electrical inductor

electrical resistor

electrical conductor

Part

Real parts

Physical elements

water pump driven by motor

source of volume flow-rate

open water tank

fluid capacitor

long pipe

fluid inductor

water-flow restriction

fluid conductor

Part

Real parts

Physical elements

weight suspended on cable via pully

source of force

mass of moving body

inertor

helical spring

pure spring

oil-filled damper

pure damper

Part

Real parts

Physical elements

combustion engine

source of torque

flywheel

rotational inertor

long flexible shaft

torsional spring

ventilator propeller

rotational damper

To demonstrate the principles of such equation formulation, let us start with the examples presented in Table 6.3 already.
(a) Constitutive relations of physical elements involved in the examples are given in Table 6.4 in three different forms where J, C, L
and G are parameters of the elements, ik is the through variable, vk is the across variable of the k-th element in each of the diagrams. Let
us choose the implicit form:

Table 6.4 Constitutive relations of the elements in Table 6.3.


Part

Physical elements

Implicit relations

Relations for i

Relations for v

through-variable source

i1 J = 0

i1 = J

capacitor or invertor

inductor or spring

conductor or damper

i4 Gv4 = 0

i4 = Gv4

(6.3
)

(b) Fig. 6.1 shows the graph of the multipole diagrams in Table 6.3. The graph nodes are identical with the nodes of the diagrams. Each
of the diagram elements is represented in the graph just by a line segment. In Fig. 6.1(a) the line segments are associated with arrows
indicating orientation of the elemental through variables (fixed to the orientation of the individual elements). According to the
postulate of continuity (6.1) for the elemental through variables at nodes M and N
(6.4
)

(c) Fig. 6.1(b) shows two independent loops in the graph. The loop a is made up by elements 1 and 2, and the loop of elements 2, 3 and
4. Arrows indicate the chosen orientation of the loops with respect to the orientation of the elemental across variables. According to the
postulate of compatibility (6.2), the elemental across variables along the loops a and b
(6.5
)

Figure 6.1 Graph of the multipole diagrams in Table 6.3 with arrows indicating orientation of elemental variables and chosen loops.
The four constitutive relations (6.3) together with the two continuity relations (6.4) and two compatibility relations (6.5) form eight
equations for eight unknown variables. In this case, the number of equations can be reduced by one after substituting i1 = J to (6.4).
In general, for a model composed from m physical elements, the described procedure, called also the tableau formulation, results in
2m differential equations to be solved to obtain 2m unknown system variables.

6.2 Reduced equations

6.2.1 Nodal formulation


6.2.2 Loop formulation

6.2.1 Nodal formulation

For practical reasons, engineers tend to decrease the number of equations to be solved simultaneously. Nodal formulation is one of the
procedures resulting in such reduced equations.
A systematic method of carrying out the nodal formulation for a model of n non-reference nodes (i.e. nodes that were not chosen as
reference nodes) is following:
Step 1. For each of the n nonreference nodes write the continuity equation (6.1).
Step 2. Express all the elemental through variables in the continuity equations in terms of the elemental across variables using the
constitutive relations of the individual elements.
Step 3. Write the elemental across variables in terms of the nodal across variables, i.e. the across variables of the nonreference nodes
relative to the reference node.
Step 4. Solve the resulting n equations for the n nodal across variables.
Step 5. Find the elemental across variables in the model as a difference of two of the nodal across variables.
Step 6. Express the through variables of the elements using their constitutive relations.

6.2.1.1 Examples
These steps can be illustrated again with the system models shown in Table 6.3.
Step 1. This step results in equations stated already in (6.4).
Step 2. Substituting the explicit constitutive relations for the elemental through variables from Table 6.4 into (6.4) gives

(6.6
)

Step 3. The elemental across variables expressed in terms of the nodal across variables vM and vN are
(6.7
)

After substituting (6.7) into (6.6) we obtain

(6.8
)

where the known variables were shifted to the right-hand sides of the equations.
Step 4. Using the nodal formulation, we have arrived at only two equations (6.8) to be solved simultaneously for the two unknown
nodal across variables vM and vN.

Step 5. The elemental across variables are expressed in terms of the nodal across variables as in (6.7).
Step 6. The elemental through variables can be evaluated using the explicit constitutive relations given in Table 6.4. Note, however,
that the nodal formulation cannot be applied directly to models with physical elements like sources of across variable, for example, as
the explicit constitutive relation for their through variables necessary in Step 2 of the formulation does not exist. Also, the integrodifferential form of the equations (6.8) does not suit to common computer computational routines for simultaneous solving of the
equations.

6.2.2 Loop formulation

An alternative formulation method for formulating equations well known from electrical-engineering textbooks is the loop formulation
(or mesh formulation for planar models). The number of equations resulting from this formulation is for a model with independent
loops. A systematic method of carrying out the loop formulation is the following.
Step 1. For each of the loops write the compatibility equation (6.2).
Step 2. Express all the elemental across variables in the compatibility equations in terms of the elemental through variables using the
constitutive relations of the individual elements.
Step 3. Write the elemental through variables in terms of the loop through variables. Step 4. Solve the resulting equations for the
loop through variables.
Step 5. Find the elemental through variables in the model as a difference of two of the loop through variables.
Step 6. Express the across variables of the elements using their constitutive relations.

6.2.2.2 Example
To illustrate these steps let us apply them to the system models shown in Table 6.3 in which the source of through variable was
replaced by the source of across variable characterized by the constitutive relation
(6.9
)

Step 1. The the compatibility equations for this example are already in (6.5).
Step 2. Substituting the explicit constitutive relations for the elemental through variables from Table 6.4 into (6.4) gives

(6.1
0)

Step 3. The elemental through variables expressed in terms of the loop through variables ia and ib are

(6.1
1)

After substituting (6.11) into (6.10) we obtain

(6.1
2)

where the known variables were shifted to the right-hand sides of the equations.
Step 4. Using the loop formulation, we have arrived at only two equations (6.12) to be solved simultaneously for the two unknown
nodal across variables ia and ib. (By adding the second equation to the first one the number of equations to be solved is reduced only to
one in this special case.)
Step 5. The elemental through variables are expressed in terms of the loop through variables as in (6.11).
Step 6. The elemental across variables can be evaluated using the explicit constitutive relations given in Table 6.4 and (6.9). Note,
however, that the loop formulation cannot be applied directly to models with sources of through variable as the constitutive relation
explicitly expressing their through variables necessary in Step 2 of the formulation does not exist. Also, the integro-differential form of
the equations (6.8) does not suit to common computer computational routines for simultaneous solving of the equations.

6.3 Extended nodal formulation

6.3.1 Principles of the method


6.3.2 Stamp submatrices of uncontrolled elements
6.3.3 Stamp matrices of controlled physical elements
6.3.4 Stamp matrices of transducers

6.3.5 Stamp matrices of blocks

6.3.1 Principles of the method

To overcome its limitations, the nodal formulation can be extended in such a way that the formulated equations include not only the
nodal across variables, but also the through variables of some of the physical elements in the model. A moderate increase in the
number of the resulting equations to be solved simultaneously is outweighed by the applicability of the extended nodal method to
models consisting of any physical elements.
Let us first demonstrate that the equations for the system models in Table 6.4 resulting from the extended nodal formulation can be
expressed directly in the first-order algebro-differential form. This form suits much better to computer computational routines than the
integro-differential form of the equations (6.8) derived using the plain nodal formulation.
The undesirable integration operation in (6.8) originated from substituting the constitutive relation of the L element for its through
variable i3 into (6.4) in Step 2 of the nodal formulation. We can avoid this by leaving i3 in (6.4) without any change, and by adding the
implicit constitutive relation of the L element from Table 6.4 as a third equation. Then,

(6.1
3)

When writing the elemental across variables in terms of the nodal across variables (6.7) we obtain the three first-order algebrodifferential equations for three unknown variables vM, vN and i3

(6.1
4)

These equations can be expressed in the matrix form

(6.1
5)

where the symbol is replaced by the operator s assuming that the product sx should be interpreted as

The left-hand-side square and right-hand-side column matrices of (6.15) can be arranged in the tabular form convenient for manual
formulation of equations

(6.1
6)

The empty entries in the tables stand for zeros.


In a similar manner, the nodal formulation can be extended towards sources of across variables or any other physical elements. In the
next paragraphs, you will learn how to form the extended nodal equations very easily following a straightforward pattern.

6.3.2 Stamp submatrices of uncontrolled elements


Table 6.5 Stamp submatrices of uncontrolled physical elements
Physical element

Without
through variable

Conductor or damper

With
through variable

va

vb

Resistor

va

vb

va

vb

va

vb

1/R

1/R

1/R

1/R

1
1

Capacitor or inertor
a

va

vb

Cs

Cs

va
Cv(0)

1
vb

R
i
1

Cs

Cs

Cv(0)

1
Cs

Inductor or spring

Cs

va

vb

Mutual inductance

Ls

i1
L1
L2

va

Cv(0)

Source of through variable

Li(0)

i2
Ms

L1i1(0)
Mi2(0)

Ms

vb

L2i2(0)
Mi2(0)
i

1
1

Source of across variable

va

vb

J
i

The stamp submatrices in Table 6.5 allow formulating the extended nodal equations as simply as imprinting stamps. The submatrix
entries should be added to the entries of the equation matrices in a tabular form similar to that in the example (6.16). The element
symbols in the table are mostly in their generic multi-physical-domain form.
The Table 6.5 gives submatrices for formulation of the equations with elemental through variables for all the physical elements listed
in the table. It gives also submatrices for nodal equation formulation without the elemental through variable for those elements for
which such submatrices exist.
The initial conditions at the time instant t = 0+ shown in the right-hand-side submatrices of some of the elements apply to the case
when the resulting equations are considered as Laplace transforms of algebro-differential equations. If this is the case, s represents the
Laplace-transform operator.
The pattern of the extended nodal formulation exploiting the stamp submatrices is easily identified.
Table 6.6 Stamp submatrices of controlled elements
Physical element

Without
through variable

Source of through variable controlled by across variable

With
through variable

vc

vd

b g

vc

vd

Source of through variable controlled by through variable

ic

ic

Source of across variable controlled by through variable

va

1
vb

ic

1
1

Source of across variable controlled by across variable

va

1
vb

r
vc

vd

1
1 1

Ideal
operational
amplifier

va

vb

1
1

Step 1. Inspect the system model and count the non-reference nodes as well as the elements the through variables of which should be
introduced as additional variables in the formulated equations. These will be

sources of across variables

inductors and springs

sources of through variables, conductors, dampers, capacitors, and inertors if their through variable is used to control an
element in the model, or if it is requested as the output variable.

Step 2. If the number of he counted elements is m and the number of the non-reference nodes is n, draw tables for the (m + n)dimensional square and column matrices. Denote the first n columns of the square table by the identifiers of the nodal across variables,
and the remaining m columns by identifiers of the computed through variables. At the same time, denote the first n rows of both tables
by the identifiers of the system model nodes and the remaining m rows by identifiers of the elements related to the computed through
variables.
Step 3. For each of the elements in the system model associate its poles a and b shown in Table 6.5 with the nodes of the system
model. Add then the entries of the element submatrices to the related entries of the resulting matrices. Ignore the entries related to the
reference node. For example, to form matrices (6.16) for the system models in Table 6.3, the poles of the involved elements should be
associated with the model nodes in the following way:

(6.17)

Try now to form matrices (6.16) using the stamp submatrices of the elements.
Table 6.7 Stamp submatrices of ideal transducers

Transducer

Without
through variable

With
through variable

Ideal transformer

va

Ideal gyrator

va

vb

vc

vb

vc

vd

vd

va

vb

vc

vd

g
g

6.3.3 Stamp matrices of controlled physical elements

The stamp-matrix approach can be easily applied also the models with controlled elements. If an element is controlled by a through
variable, this controlling through variable should be introduced as the additional variable in the formulated equations.

Table 6.6 gives stamp matrices of controlled sources. The controlling through variables are denoted as iC, while iC stands for the
controlling across variables between system model nodes c and d.

6.3.4 Stamp matrices of transducers

Table 6.7 demonstrates that the extended nodal method can be easily applied also to systems containing such four-pole elements like an
ideal transformer, an ideal gyrator, or an ideal operational amplifier.

6.3.5 Stamp matrices of blocks

Table 6.8 Stamp submatrices of basic blocks


Block

Stamp submatrix
va

signal source

va

vb

scalor

va

vb

vc

comparator

va

vb

va(0)

integrator

va

vb

vb(0)

differentiator
vA
a

vB

1
1

transfer block

x1

x2

x3

b0

b1

b2 sb3

b3x3(0)

a0

a1

a2 s

x3(0)

x1(0)

x2(0)

The above given procedure for equation formulation can be applied also to system models represented by block diagrams as well as by
multipoles combined with blocks. Table 6.8 gives the stamp matrices for basic blocks.
6.4 Block diagrams

6.4.1 Construction of block diagrams


6.4.2 Block diagrams vs. multipoles

6.4.1 Construction of block diagrams

Though block diagrams might look similar to multipole diagrams at the first sight, their principle is very different. We shall
demonstrate the procedure required for the block-diagram construction using again the system examples given in Table 6.3.
Step 1: Formulation of a set of equations.
Any of the school-book pencil-and-paper procedures yielding consistent equations characterizing the system model can be used for this
purpose.
Let us utilize for our demonstration the implicit equations (6.3), (6.4) and (6.5) formulated in the previous section.
Step 2: Decreasing the order of the differential equations.
Any n-th order differential equation can be converted to n first-order equations by substitution for the higher-order derivatives.

There are no higher-order equations among those considered in our demonstration, however.
Step 3: Conversion of equations into explicit form.
Such a conversion should result in a set of equations each of which expresses explicitly one of the unknown variables.
Table 6.4 shows the constitutive relations of the physical elements from our example in two different explicit forms (except the
through-variable source for which only one explicit form exists). So, we are supposed to choose one of the explicit expressions for
each of the elements and, at the same time, to rearrange the relations (6.4) and (6.5) into the explicit form in such a way that there is
just one relation for each variable. Therefore, we are faced with the so called causality problem.
For computational reasons numerical integration is preferred to numerical differentiation in block-diagram oriented packages. Let us
just form the first three explicit relations as

(6.1
8)

Once we made this choice, it can be proven that the rest of the explicit relation must be formed in the following way:

(6.1
9)

For example, if we had chosen the relation i4 = Gv4 for the fourth element, we would not be able to solve the problem. For some system
configurations, even some of the integral relations must be replaced by differential ones to solve the causality problem.

Figure 6.2 Block diagram representing the systems in Table 6.3.


Step 5: Setting up the block diagram.
The resulting block diagram is shown in Fig. 6.2 using the DYNAST graphical notation. Each of the blocks in the diagram represents
just one of the explicit relations from (6.18) and (6.19) (no blocks are needed for the trivial relations for i4 and v1).

6.4.2 Block diagrams vs. multipoles


Let us summarize the experience learnt from this simple example:

If we want to represent energy interactions of a real subsystem by a block, for most of the subsystems we must have at our
disposal two different blocks at least, each of them representing the same constitutive relation, but expressing explicitly
different variables. In the case of the multipole approach, only one model is sufficient for each subsystem, and this model is
reusable in any consistent system configuration.

In the block-diagram model of a complete system, in addition to the blocks representing the individual subsystems, there must
be also blocks respecting the postulates of continuity and compatibility. The latter blocks are not reusable as they suit just to

one specific system configuration. For the multipole models the postulate-respecting relations can be formed automatically and
no additional blocks are needed.

To choose which block should model a particular subsystem, and to form blocks respecting the postulates, we have to solve the
causality problem. In the multipole approach such a problem is not encountered at all.

The manual procedure for setting up a block diagram for the physical-level model of a real-life system, usually nonlinear, is much
more involved than could be shown in this example, of course. However, there is also a big difference between DYNAST and the
common block-diagram oriented packages in the way of solving equations.
Once a block diagram is submitted to DYNAST, a set of implicit algebro-differential equations is formed, and these are solved
simultaneously as it is the case of the equations for multipole models. The common block-diagram oriented programs usually do not
solve any equations in reality. They would not treat the relations (6.18) and (6.19) as equations, but as assignments.
The variables are processed successively by one block only at a time at discrete time steps. The process starts from blocks with no
input like the block specifying
in Fig. 6.2. But this is not so simple in the case of feedback loops in the block diagram.
For example, the summing block for the output variable
cannot specify the correct value of this variable at the
end of the first time step, but only after a number of time steps and iterations. This problem is even more difficult if there is no
integrator delaying the variable propagation in the feedback loop by one step. Solving such algebraic loop problems requires special
techniques.
In many cases, the blocks in a block diagram must be rearranged before the simulation starts as their order is critical to the validity of
the computed results. There are, however, system models characterized by systems of differential equations the order of which is timevariable. Solving such an change-of-order problem requires reconfiguration of the block diagram not only prior, but also during the
simulation.
As all the equations are solved simultaneously in DYNAST, there are no algebraic-loop problem. As, in addition, the equations are
solved in DYNAST directly in their natural implicit algebro-differential form, there is no change-of-order problem. Not only this, the
related computational procedures for such equations are in principle more efficient and robust then those implemented in the common
block-diagram packages. Obviously, developing software for automated construction of physical-level system models as a preprocessor for a common block-diagram oriented package is not the optimal way of coping with the problem.

6.5 Lagranges equations of multipoles

Multipole models of subsystems can be characterized by constitutive relations formed using different approaches. Lagranges
equations of second kind combined with undetermined multipliers can be exploited to form the constitutive relations using the
following procedure.
Step 1. Identify the subsystem energy entries and mass centers of the bodies within the subsystem.
Step 2. Determine the degree of freedom i of the subsystem bodies.
Step 3. Choose

coordinates sj, j = 1 , 2 , ... , n, to characterize motions of the entries and of the body mass centers.

Step 4. Form r = n i relations between the coordinates implied by the geometric constraints
(6.2
0)

Step 5. Differentiate (6.20) to obtain n i relations for the subsystem geometric compatibility
(6.2
1)

Step 6. Consider the free diagram, i.e., consider the disconnected from a system. Show all the external forces the system exerts on the
to hold it in equilibrium.

Step 7. Formulate n Lagranges equations in the form


(6.2
2)

where , , and is the kinetic, potential, and dissipative energy of the. Qj is a generalized external force associated with the
generalized coordinate sj, whereas k is the Lagranges multiplier representing the generalized reaction force associated with the
coordinate sk.
Step 8. By excluding the r multipliers k from (6.22) obtain n r equations for quasi-static balance of inertial, potential and dissipative
forces.

6.5.1 Example: Body with two revolute joints

Figure 6.3 (a) Body with two revolute joints, (b) its multipole model.

The subsystem shown in Fig. 6.3a is a body constrained to a planar motion with two revolute joints denoted by A and B. O denotes the
mass center of the body. Each of the joints is represented by two poles in the multipole model of the body given in Fig. 6.3b. Poles Ax
and Bx represent translational motions of the joints along the x-axis whereas the poles Ay and By along the y-axis. Pole O corresponds
to the rotational motion of the body about the mass center O.
The body shown in Fig. 6.3a is of i = 3 degrees of freedom. Let us consider n = 7 coordinates: xA , yA , xB , yB , xO , yO, and . Then the r
= n i = 4 geometric constraints can be expressed as:

By differentiating the constrains we obtain the first part of the constitutive relations:
(6.2
3)

The body kinetic and potential energies are

where JO is the moment of inertia of the body with respect to its center of mass.
Substitution into (6.22) yields for the individual coordinates:

Finally, after excluding the multipliers, we are arriving at the equations for the force dynamic equilibrium forming the second part of
the constitutive relations for the multipole model of the body with two revolute joints

6.5.2 Example: Flexible crank-and-slide mechanism

Figure 6.4 (a) Slider-crank mechanism with a flexible connecting rod, (b) its multipole model.
Fig. 6.4a shows a slider-crank mechanism which is considered as massless, but its connecting rod is assumed to be flexible with
dissipative losses. In the corresponding multipole model given in Fig. 6.4b, the rotational motion of the crank is represented by the
pole B whereas the translational motion of the slider represents the pole Ax.

The mechanism shown in Fig. 6.4a is of i = 2 degrees of freedom. Let us consider n = 4 coordinates: xA , A , B and s. Then the r = n
i = 2 geometric constraints can be expressed as:

By differentiating the constrains with respect to time we obtain:


(6.2
5)

The potential and dissipative energies are in this case

where k and b is the stiffness and damping factor of the connecting rod, respectively.
Substitution into (6.22) yields for the individual coordinates:

Finally, after excluding the multipliers, we can complete the constitutive relations by the equations

Module 7
Formulation of transfer functions
Module units

7.1

Formulation of system-response transforms

7.2

Determinant evaluation

7.3

Formulation of system transfer functions

7.4

Solved examples

Module overview.
This module will introduce you to formulation of Laplace transforms of systems responses based on transforms system equations
formulated in an arbitrary way. In the second unit you will learn how to form arbitrary transform functions from the system equations.

Module objectives. When you have completed this module you should be able to:
1. Understand transforms of system responses and transfer functions in relation to system equations.
2. Formulate transforms of system responses from equations of linear system models.
3. Formulate system transfer functions from equations of linear system models.
Module prerequisites. This course module on Formulation of system equations.

7.1 Formulation of system-response transforms

Let us consider a linear time-invariant dynamic system described in the Laplace transform domain by a set of n algebraic equations.
Such equations, resulting e.g. from the extended nodal formulation, are of the general form
(7.1
)

where A(s) is a given n-dimensional square matrix, X(s) is the n-dimensional column vector of the system variables to be determined,
B(s) is the given
matrix, U(s) is the p-dimensional column vector which represents the system excitation by independent sources,
and C(0+) is the given n-dimensional column vector representing the system initial conditions related to the energy accumulated in the
system.
Obviously, the Laplace transform X(s) of the unknown vector x(t) can be written immediately in terms of the inverse matrix A1(s) as
(7.2
)

After the inverse Laplace transformation is applied to (7.2), the vector of system responses x(t) can be expressed as the sum of two
response components

where

is the response of the system with zero initial conditions to an independent excitation, and

is the response of the system with zero excitation from independent sources to system initial conditions.
Thus, the complete solution of (7.1) involves no more than the determination of the inverse of a given matrix and the inverse Laplace
transformation. Nevertheless, to calculate the inverse of a matrix, say, anything more than a
matrix, is not a simple job without the
help of a computer.
However, if we are required to find only one or two of the unknowns in (7.1), we need not compute the inverse of the matrix A. Using
the Cramers rule, the i-th component of X(s) in (7.1) can be expressed as
(7.3
)

assuming that the determinant (s) of the matrix A(s) is nonzero. i(s) is obtained form (s) by replacing the i-th column of (s) by
the right-hand-side of (7.1).

7.2 Determinant evaluation

Determinant of a

matrix A = [a] is defined as = a. Determinant of a

matrix

is

Determinants of larger matrices can be computed by means of the Laplace expansion. In general, expansion of the determinant of an
n-dimensional square matrix A along its k-th row, resp. along its l-th column, yields
(7.4
)

where akl is the element in the k-th row and l-th column of the determinant, and kl is the (n 1)-st order determinant obtained from the
determinant by deleting the k-th row and l-th column of .
For example, the determinant of the order n = 3 associated with the

matrix

expanded along its first row

You may wish to show that expanding along another row or column gives the same result. This method is valid for any n. Determinant
evaluation using the Laplace expansion, however, is not suitable for the analysis of large systems as it requires (n 1)n!
multiplications. While for n = 3 the number of multiplications is 12, for n = 4 it becomes 72 already.
Fortunately, the system matrices are usually rather sparse so that the formulas (7.4) give us the chance to take advantage of this matrix
property. In general: the larger is the analyzed system, the higher is the percentage of zero-value elements in its matrix A. By starting
the determinant expansion from the row or column with the least number of nonzero elements we can reduce the amount of
computations considerably.
7.3 Formulation of system transfer functions

In many design situations, rather than system responses, we need to compute a system transfer function (or a matrix of such functions).
Let us thus assume, that a system transfer function Hij(s) of a system described by (7.1) is defined as
(7.5
)

where the dependent variable Xi(s), which is the i-th element of the vector X(s), is taken as the system output, and the independent
excitation variable Uj(s), being the j-th component of the vector U(s), is considered as the system input. (According to the transfer
function definition C(0+) = 0.)
Utilizing the Cramers rule (7.3), we can express the transfer function (7.5) as
(7.6
)

where

is the determinant of the matrix A(s) the i-th column of which has been replaced by the j-th column of the v B(s).

7.4 Solved examples

Mechanical translation system. Consider the spring-mass-dashpot system shown in Fig. 7.1. Let us obtain the transfer function of the
system
(7.7
)

by assuming that the force F(t) is the input and the displacement y(t) of the mass is the output.

Figure 7.1 Mechanical translational system.


The multipole model of the system is given in Fig. 7.1a. The node A represents the vertical motion of the mass. The polarity of the
input-force source has to comply with the chosen orientation of the velocity
. In this simple case only one equation is needed to
describe the dynamics of the given system. This is the equation in Fig. 7.1b are
F
A

b + sm + k/s

The transfer-function H(s) can then be readily expressed as


(7.8
)

Mechanical rotational system. Consider the system shown in Fig. 7.2a. The system consists of a load inertia and a viscous-friction
damper driven by a rotational motor by means of a flexible shaft. The motor is modelled as it is seen from the load.
System parameters:

m = motor torque applied to the system


bm = motor damping (corresponding to the slope of its loading characteristics)
k = shaft torsional stiffness

Figure 7.2 Mechanical rotational system.


Jl = moment of inertia of the load
bl = load viscous-friction
l = load angular velocity
Let us obtain the transfer function of the system
(7.9
)

The multipole diagram model of system is given in Fig. 7.2b. The nodes A and B represent the rotational motions of the motor and the
load, respectively. The equations describing the dynamics of the system are

bm + k/s

k/s

k/s

bl + sJ + k/s

The transfer-function H(s) can then be expressed as


(7.1
0)

where
(7.1
1)

Thus
(7.1
2)

L-R-C circuit. Consider the electrical circuit shown in Fig. 7.3a. The circuit consists of an inductor L, a resistor R, and capacitor C.

Figure 7.3 L-R-C circuit


If ei is assumed to be the input and v0 the output, then the transfer function of the system is assumed to be
(7.1
3)

The multipole diagram of the system is given in Fig. 7.3 b. It differs from the actual circuit diagram by the voltage source representing
the input excitation ei.
When formulating equations describing the system, we shall first take into consideration all the three nodes shown in the diagram:
v1

v2

v3

ei

3
ei

G
1

G + sC
1

The transfer function H(s) can then be expressed as


(7.1
6)

In this case we can considerably simplify the problem by formulating just two equations, one for the node 3 and the second one for the
series combination of the elements ei L R. Then we shall obtain

In this case

v3

sC

ei L R

sL R

ei

(7.1
7)

The resulting transfer-function H(s) is identical to (7.16).


Seismograph. Fig. 7.4a shows a schema of seismograph. A seismograph indicates the displacement of its case relative to the inertial
reference. It consists of a mass m fixed to the seismograph case by a spring of stiffness k , b is a linear damping between the mass and
the case.
Let us define:
xi = displacement of the seismograph case relative to the inertial reference (the absolute frame)
x0 = displacement of the mass m within the seismograph relative to its case
y = x0 xi = displacement of the mass m relative to the case.
Note that y is the variable we can actually measure.Since gravity produces a steady spring deflection, we measure the displacement x0
of mass m from its static equilibrium position.
Considering xi as input an y as output, the transfer function is
(7.1
8)

Figure 7.4 Seismograph


The multipole model of the seismograph is given in Fig. 7.4b. It is advantageous to take there as a reference node the node C
corresponding to the case instead of the node F corresponding to the actual inertial reference. The transfer function can than be
expressed as
(7.1
9)

Exploiting the fact that the velocity between nodes F and C is constrained by the input velocity source vi, we can write the matrix-form
equations for the multipole diagram in the following way
vMC

vFC

b + sm + k/s

sm

vi

The determinants are in this case


(7.2
0)

From this
(7.2
1)

Note that
(7.2
2)

Thus for low frequency inputs xi the mass m follows the case up and down. However, if the input variable has a sufficiently high
frequency, then the mass m remains almost fixed with respect to the inertial reference and the motion of the case indicates the relative
motion between the case and the mass. For this reason the undamped natural frequency
Liquid-level system. Consider the system of two interacting tanks shown in Fig. 7.5a.

is made small in seismographs.

Figure 7.5 Interacting tanks.


The fluid capacitance of the tanks is
(7.2
3)

where A1 , A2[m2] are horizontal areas of the tanks, [kg.m3] is the mass density of the liquid, and g[m.s2] is the acceleration due to
gravity. The volume fluid flow into the first tank through the control valve is Qi. The tanks are interconnected by a short pipe with a
valve representing a restriction of fluid R1. The fluid resistance of the second tank outlet is R2.
The task is to compute the transfer function between the intake flow Qi and the outlet flow Q2 which can be expressed as
(7.2
4)

where p2 is the liquid pressure at the bottom of the second tank.

Based on the assumption that the system is either linear (for the laminar flow in the restriction) or linearized (for the turbulent flow), it
can be modelled by the multipole diagram given in Fig. 7.5 (the fluid conductances G1 = 1/R1, G2 = 1/R2).
The port diagram equations are then
p1

p2

Qi

G1 + sC1

G1

G1

G1 + G2 + sC2

The relevant determinants are


(7.2
5)

From there
(7.2
6)

Multivariable mechanical system. Consider the system shown in Fig. 7.6a. We assume that the system, which is initially at rest, has
two inputs F1(t) and F2(t) and two outputs x1(t) and x2(t). In this problem we would like to demonstrate formulation of a multivariable
system transfer-function description

Figure 7.6 Multivariable mechanical system.


(7.2
7)

where the transfer-function matrix


(7.2
8)

The equations describing the multipole diagram model of the system given in Fig. 7.6b are
F1
1

b1 + sm1 + k1/s

b1

b1

b1 + sm2 + k2/s

F2

1
1

The transfer-function matrix H(s) can then be expressed as


(7.2
9)

where
(7.3
0)

and

(7.3
1)

Clearly, the time responses x1(t) and x2(t) are given by

assuming zero initial conditions.


Show, that the equivalent of inertia and friction of the gear train referred to shaft 1 are given by

Similarly, the equivalent moment of inertia and friction of the gear train referred to the load shaft are

The relation between J1eq and J3eq is thus


(7.3
8)

and that between b1eq and b3eq is


(7.3
9)

Note, that if

and

for a speed-reducing gear train, then the effect of J2, J3, b2 and b3 referred to shaft 1 is negligible.

Rotary-to-linear motion conversion. In motion control problems, it is often necessary to convert rotational motion into a translational
one. For instance, a load may be controlled to move along a straight line through a rotary motor and screw assembly, such as that
shown in Fig. 7.7a.

Figure 7.7b shows a similar situation in which a rack and pinion is used as the mechanical linkage. Another common system in motion
control is the control of a mass through a pulley by a rotary prime mover, such as that shown in Fig. 7.7c.

Figure 7.7 Rotary-to-linear motion conversion.


The system shown in Figs. 7.7a, 7.7b and 7.7c are all represented by a simple port diagram in Fig. 7.7d. Elements m, Jm and bm are
modeling the drive motor there, and the elements , and model the mechanical load of the systems. The ideal transformer models
the actual motion conversion.

In case of the system shown in Fig. 7.7a, the transformer ratio n corresponds to the lead of the screw, which is defined as the linear
distance which the load travels per revolution of the screw. In the other two systems n = r where r is the radius of either the pinion or
of the pulley.
The equations describing the dynamics of the systems are following:
M
M

vL

b + sJm

L
transformer

1
n

The determinant of the system matrix is


(7.4
0)

Using this, we can easily derive the following transfer functions

From the second transfer function we can conclude that the equivalent torque e, inertia Je and damping be of the load that the motor
sees is

Thus any of the three rotary-to-linear conversion systems can be represented from the point of view of loading the drive motor just by
the elements e, Je and be as shown in Fig. 7.7e.
Note also, that the third transfer function becomes

for bm = Jm = 0 as could be expected.


From the last transfer function it is obvious, that the load sees the motor as shown in Fig. 7.7f where the elements Fe, me and be are

Rotary-to-rotary motion conversion. Consider the system shown in Fig. 7.8a. In this system, a load is driven by a motor through the
gear train. Assuming that the stiffness of the shafts of the gear train is infinite (there is neither backslash nor elastic deformation) and

that the number of teeth on each gear is proportional to the radius of the gear, obtain the equivalent moment of inertia and equivalent
friction referred to the motor shaft and referred to the load shaft.

Figure 7.8 Gear train system.


In Fig. 7.8a, the numbers of teeth on gears 1, 2, 3 and 4 are N1 , N2 , N3 and N4, respectively. The angular displacements of shafts are
1 , 2 , 3 and 4, respectively. Thus, 2/1 = N1/N2 = n21 and 4/3 = N4/N3 = n43. The moment of inertia and viscous friction of
each gear train component are denoted by J1b1 , J2b2 , J3b3, and J4b4, respectively. J1b1 include the moment of inertia and friction of the
motor, J4b4 include the moment of inertia and friction of the load.
The multipole model of the gear train system is given in Fig. 7.8b. The equations describing the multipole diagram model of the
system given in Fig. 7.8 are

2
1

b1 + sJ1

n21
b2 + sJ2

3
T12
T34

1
b3 + sJ3

n21

1
n43

1
n43
1

You might also like