Вы находитесь на странице: 1из 66

Third Year Dynamics Lecture Notes

Michael Zaiser
Contents
1 Introduction 1
1.1 Prerequisites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Course outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.3 Use of these notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.4 Basic concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2 The theory of systems with one degree of freedom 5
2.1 Lumping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2 Free vibration of a SDF system with no damping . . . . . . . . . . . . . . . . . . 6
2.2.1 Equation of motion approach . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2.2 Energy approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3 Free vibration of a SDF system with damping . . . . . . . . . . . . . . . . . . . . 10
2.3.1 Viscous damping and solution of the damped motion equation . . . . . . . 10
2.3.2 Canonical form of the damped free vibration equation . . . . . . . . . . . 14
2.3.3 Rotational vibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.4 Periodically forced vibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.4.1 Steady-state response of a periodically forced system . . . . . . . . . . . . 19
2.4.2 Force transmission and vibration damping . . . . . . . . . . . . . . . . . . 22
i
2.4.3 Moving boundary condition and amplitude transmission . . . . . . . . . . 23
2.5 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.5.1 Accelerometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.5.2 Rotating out of balance rotor . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.5.3 Out of balance in reciprocating internal combustion engines . . . . . . . . 29
2.6 Non-periodic forcing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.6.1 Shock damping and shock response spectrum . . . . . . . . . . . . . . . . 39
3 The theory of systems with many degrees of freedom 42
3.1 Free vibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.2 Eigenvalue problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.3 Forced vibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.4 Orthogonal modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.5 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.5.1 Shaft whirling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.5.2 Beating . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.5.3 Anti-resonance and vibration absorbers . . . . . . . . . . . . . . . . . . . 58
4 Self-excited vibration 61
4.1 Positive Feedback . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.2 Stick-slip motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
ii
1 Introduction
1.1 Prerequisites
You have already studied some freely vibrating systems last year, and we will be building on this
knowledge. You will need to draw on the material you studied both in dynamics, mathematics and
solid mechanics. As the course starts, make sure you know how to draw free body diagrams (FBDs)
and extract the equations of motion from these diagrams. You have studied the theory of second
order ordinary differential equations with constant coefcients in second year maths. We will make
a lot of use of the results, so revise them! I will be going over some of the material in the rst few
lectures, but it will only be to refresh your memory and to dene my own notations, so dont rely on
learning it for the rst time here. As a test, before you start the course you should be able to solve
x +5 x +4x = exp(3t) (1)
with initial conditions
x(0) = 0 , x(0) = 1 . (2)
Remember you will need to nd the two complementary functions and the particular integral and
then use the initial conditions to determine the constants in the general solution. Sound familiar?
Could you solve the same equation if the right hand side was sin(3t)? You will also need to know
how to use matrices to represent systems of equations and know the meaning and use of eigenvalues
and eigenvectors. To deal with the problems in tutorial and exam questions, you also should recall
what you have learned about rotational motion, moments of inertia and beam exure.
1.2 Course outline
The focus of the course is dynamic vibration. You can nd more details and a lecture by lecture
breakdown in the syllabus, available from the dynamics web page. We will study the theory of
vibration of single and multiple degree of freedom systems with and without forcing. With this
theoretical work, we will be able to look at a range of applications. These will include shaft whirling,
vibration dampers, balancing in internal combustion engines and vibration measuring devices.
We will look rst at single degree of freedom (SDF) systems in some detail because they exhibit
many of the features of more complicated systems while remaining relatively easy to analyze math-
ematically. After reviewing free body diagrams and clarifying notation, we will consider a freely
vibrating SDF system with and without damping. The equation of motion for the system follows
from the free body diagram. It is a second order, linear, homogeneous ordinary differential equation
with constant coefcients and it can be tackled by methods which will already be familiar from pre-
vious courses. We will look at the physical signicance of the solution and see how the parameters
of the system determine its vibrational characteristics. Having understood the behaviour of freely
1
vibrating systems, we will introduce a forcing term and consider its effects. Some new mathemat-
ical techniques will be introduced here. We will rst look at systems forced by a periodic force
and then at the more difcult general case. At this point we will develop a general solution for all
SDF systems, free or forced and damped or undamped, and discuss the different properties of each
system and the methods used to analyse them.
The second part of the module will address multiple degree of freedom (MDF) systems. We will
see that the problem is similar to an SDF system but with some added complications. We nd that
the mathematical analysis is similar in SDF and MDF systems if we replace the mass and stiffness
constants with matrices. We will look at how to obtain solutions of the matrix equations and what
these solutions tell us about the behaviour of the system. We will see that in many cases we can
make a coordinate transformation that will allow us to solve the MDF system as a series of SDF
systems. The equations for MDF systems are often very hard to solve. In practice they are usually
solved numerically, and we will look at some of the methods that are used to do this.
Along with the mathematical theory, we will look at some applications to engineering systems. We
will nd in most instances that we must simplify the true situation but that the answers we get from
the theory agree quite well with experiments. We will look at shaft whirling, IC engine balanc-
ing, rotating out-of-balance systems, vibration absorbers, anti-resonance devices, accelerometers,
torsional and beam systems. Numerical methods are becoming the dominant method of analysing
vibrating systems and we will spend some time using MATLAB to explore how the methods work
and what we need to be able to do to use them. It is a common misconception that numerical meth-
ods are easy to use. We will see that it is indeed almost too easy to use numerical methods to get
answers, but rather more difcult to use them properly and get valid answers.
At the end of the course we will look at what we have achieved and hopefully conclude that we can
analyse many interesting systems that are important in engineering. We will also discuss what we
have not yet studied and see the limitations of our theories.
1.3 Use of these notes
I suggest that you will nd the material easier to follow if you look at the relevant sections of these
notes before the lectures. Do the exercises in the notes as we go along; they will help your under-
standing. You need to be active in your study by trying out the techniques on problems yourself.
Passively reading the notes will not get you very far.
It should really go without saying that these notes are in no way a replacement for the lectures, and
not only because they are incomplete. The notes complement the lectures and relieve you of the
tedious, error-prone task of copying chunks of algebra from the board.
2
1.4 Basic concepts
We will be building on the work from previous courses. You should already be familiar with the
basic ideas of freely-vibrating systems. In particular, you ought to know the meaning of period and
frequency of vibration, inertial forces, spring stiffness, moments of inertia, beam bending, comple-
mentary functions and particular integrals. You should also be able to draw free body diagrams
correctly. We will revise free body diagrams before starting the course proper.
A free body diagram shows the real and inertial forces that act on each mass in a system. Once a
free body diagram has been constructed, the equation of motion can often be written down. When
constructing the free body diagram, it is important to bear in mind the direction of each force in the
system.
x
1
x
2
k
1
m
1
k
2
m
2
k
3
Figure 1: An illustrative two-degree of freedom system. The masses are free to move in the hori-
zontal direction only. When x
1
= 0 and x
2
= 0 the system is in static equilibrium
The best way to see how this works is to consider an example. In Figure 1 the two objects with
masses m
1
and m
2
are free to move in the horizontal direction. However, they are not free to move
in the vertical direction, an example of a constraint. The masses can move independently, so two
coordinates are needed to specify their position. The system therefore has two degrees of freedom.
The three springs in Figure 1 have stiffness constants k
1
, k
2
and k
3
and like all springs we will
consider they are linear. If a linear spring is extended by a distance x, it exerts a force kx where
k is its stiffness. From this information, we can draw the free body diagram for the system. As a
reference conguration we always chose the static equilibrium conguration of the system. Figure 2
shows the system after the masses have been moved by (arbitrary) distances of x
1
and x
2
out of this
conguration. If the distance between the masses in the equilibrium conguration is l, the distance
between them after they have moved is l x
1
+x
2
. The extension of the central spring is therefore
x
2
x
1
. The forces due to the springs can now be deduced. Figure 3 shows the forces acting on each
mass, including the inertial forces. Note that the nal spring is compressed when x
2
is positive and
so the force must be directed to the left. We choose to show this by drawing the arrow to the right
and including a minus sign. Equivalently, we could have kept the force as positive and directed the
arrow to the left.
3
x
1
x
2
l
l-x
1
x
2
Figure 2: Distances when the masses have moved.
k
1
x
1
k
2
(x
2
-x
1
)

-k
3
x
2
1 1
x m
2 2
x m
Figure 3: The free-body diagram for the illustrative system.
The direction of the inertial forces is important, and can be determined by considering the effect
of the inertia of the mass. If x
1
is positive, as it is shown in Figure 2, then the mass m
1
has been
moved to the right. If we accelerate a mass in the direction of x
1
to the right, its inertia acts to
oppose this and so the inertial force is directed to the left. The same argument applies to the mass
m
2
. Hence, the direction of the inertial forces must always be taken opposite to the direction of the
respective coordinate. The equations of motion for the two mass system can now be written down
using DAlemberts principle that the sum of the forces on each body must be zero in dynamic
equilibrium. We obtain one equation for each mass, and they are given by
m
1
x
1
+k
1
x
1
k
2
(x
2
x
1
) = 0 ,
m
2
x
2
+k
2
(x
2
x
1
) +k
3
x
2
= 0 . (3)
Exercise: 1
It is a useful exercise at this point to redraw the system of Figure 1 with a different
coordinate system. Label the system with coordinates x
3
and x
4
so that x
3
is just
the same as x
1
but x
4
runs in the opposite direction to x
2
. Draw the free body
diagram and deduce the equations of motion. When you have done this, substitute
x
3
= x
1
and x
4
= x
2
and check you have the same equation of motion as derived
above. Tip: You may nd you need to multiply one of the equations by -1 to make
it the same as equation (3).
4
What can we conclude from this brief exercise? Minus signs can be a headache when we come to
gure out the directions of forces, but actually the mathematics takes care of all this for us! If we
set up a coordinate system so that things work properly for positive displacements, everything will
also work for negative displacements. Therefore, when setting up a FBD, just specify a direction for
the displacement of a body (and stick to it!), and draw in the correctly directed spring force which
results from this displacement. If in reality the body moves the other way, it doesnt matter because
the maths will sort out the appropriate sign on the force.
2 The theory of systems with one degree of freedom
Systems with one degree of freedom are very important and will be the starting point for our study
of vibration. We will nd that there are many interesting engineering problems that can be modelled
well by single degree of freedom (SDF) systems. Other more complicated systems will need extra
degrees of freedom, but we will discover that many of the ideas and techniques we develop for
SDF systems will be easily extendible to many degree of freedom (MDF) systems, although the
mathematical effort needed to solve them will be greater.
2.1 Lumping
Before we start to analyse a SDF system, let us consider for a moment what we are really thinking
about when we draw our initial diagram for a system. In reality, we can never cover all the compli-
cated aspects of a real system. For instance, consider a mass supported by a spring in the absence
of gravity. This appears to be a very simple dynamic system to study, and we might draw a diagram
like that shown in Figure 4.
Remember there is no gravity so the equilibrium position of the mass is at the unextended length
of the string. If we look at the details, the dynamic behavior of this system may contain many
complexities: The mass of the object supported by the spring is in reality distributed over the object
rather than concentrated at a point. The stiffness of the spring is similarly spatially extended. The
spring has a nite mass and its response is unlikely to be exactly linear. However, if we disregard
these complexities and treat the system as a point mass supported by a massless spring, the behavior
of our model system and the true physical system turn out to be very similar. We have captured the
essential features of the system in our simple model, without including all the complex details that
have only small effects.
The process whereby we consider the mass to be concentrated at a point rather than distributed
in space is called lumping and we refer to the model as a lumped parameter model. Lumping is
a very powerful idea because it greatly simplies the mathematical treatment of the system while
managing to retain the essentials. Masses and springs can both be lumped, i.e. replaced by a single
5
m
x


k
Figure 4: The simplest SDF system.
element. We will see later that another major component of our modelled systems, the damping,
can also be lumped. In all cases, the essential point to bear in mind is that we are modeling the real
physical situation. We know that the mass is not concentrated at a point in reality, but that we can
model it as being so and obtain a good approximation of the real situation.
2.2 Free vibration of a SDF system with no damping
We now start our mathematical analysis of the system shown in Figure 4. We will start off in the
normal way by drawing the free body diagram, deriving the equation of motion and nding solutions
to this equation. We will also look at the energy of the system and see how this too can be used to
derive an equation of motion. Of course, the two methods give the same results.
2.2.1 Equation of motion approach
This is the way we will tackle most problems in the course. First we draw the free body diagram as
shown in Figure 5.
6
m
x


kx
x m
Figure 5: The free body diagram for the system of Figure 4.
Now we can write down the equation of motion
m x +kx = 0 . (4)
Exercise: 2
Consider the same system but with the effects of gravity included. Draw the free
body diagram and write down the equation of motion. Show that by appropriately
shifting the x coordinate one obtains the same equation of motion (4) as in the
gravity-free case. (Chose as x = 0 the static equilibrium conguration of the
system with gravity)
Before we try to solve this equation, we will classify it. The equation is second order in time because
the inertial term involves a second order derivative. The equation is linear, since the coefcients do
not depend on x, and it has constant coefcients because m and k do not depend on t. It is also
homogeneous because there is no term independent of x. In full then, the equation is a second order
homogeneous ordinary differential equation with constant coefcients.
To solve this equation, we need to nd two complementary functions. We will not need a particular
integral because the equation is homogeneous. The equation is second order, so we expect to obtain
two complementary functions. The sum of these two functions is the general solution. We can see
by inspection that either of
x = sin(t) ,
7
x = cos(t) , (5)
is a solution of the equation. The constant must be chosen so that the m and k constants are
accounted for correctly. The value needed to do this is =
_
k/m.
Exercise: 3
Verify the value given for . To do this, consider sin(t) rst. Compute the
second derivative with respect to time and substitute into equation (4). From the
resulting equation you can calculate the relation between and k and m. Repeat
the calculation for the cosine solution.
From the complementary functions, we can now construct the general solution. It is
x = Asin(t) +Bcos(t) . (6)
It is easier to see what is going on if we rewrite the general solution for x in a different form. We
can write the solution as
x =Csin(t +) . (7)
where the constants are given by C =

A
2
+B
2
and tan = B/A.
Exercise: 4
Show this is the case. Hint: start with x =Csin(t +) and expand the sine using
the formula sin(X +Y) = sin(X)cos(Y) +sin(Y)cos(X).
From this expression we can see more clearly what is happening. The constant C is the amplitude
of the vibration and is the phase. These two constants are undetermined in the general solution
and must be calculated from the initial conditions.
Lets now put some numbers into the formula, specify initial conditions, and see what the solution
looks like. Suppose m = 1 kg and k = 25 N/m. If we set the mass off at time zero at its equilibrium
position with a velocity of 1 cm/s, what does the subsequent motion look like? We calculate rst
and nd = 5 rad/s. The motion of the mass is therefore given by
x =Csin(5t/s +) . (8)
From the initial conditions, we know that x(0) = 0 which implies = 0. The velocity at any time t
can be found by differentiating the above equation to get
v(t) = x(t) = [5/s]Ccos(5t/s +) (9)
and we now know = 0 so at time t = 0 the velocity is given by
v(0) = [5/s]C . (10)
The initial conditions tell us that v(0) = 10
2
m/s so C = 210
3
m. The nal solution then is
x(t) = [210
3
m] sin(5t/s) . (11)
This solution is shown in Figure 6 for the rst three seconds of motion.
8
0 1 2 3
-2
0
2


d
i
s
p
l
a
c
e
m
e
n
t

[
m
m
]
time[s]
Figure 6: The displacement of the mass in the system of Figure 4 as a function of time. The
parameters and initial conditions are given in the text.
2.2.2 Energy approach
We can tackle the same problem discussed in Section 2.2.1 using an alternative approach based on
the energy of the system. We will nd that we can actually derive the equation of motion once we
know the energy of the system, so all the results from the previous section can be re-obtained. The
advantage is that we dont need to consider the forces in the problem, which in some cases can
be difcult. On the other hand, the energy approach as discussed here works only if there is no
damping.
Consider the energy stored in the spring. The force exerted by a spring is F = kx where x is the
extension from the equilibrium length and k is the stiffness. The work done in extending the spring
by a length x is
W =
_
Fds =
kx
2
2
. (12)
The energy stored in the spring, its potential energy, is the negative of the work done on the spring
and we will call this U(x). We have
U(x) =
kx
2
2
. (13)
9
The kinetic energy of the mass we will call T. It is given by
T =
m x
2
2
, (14)
so the total energy of the system E =U +T is
E =
m x
2
+kx
2
2
. (15)
The principle of conservation of energy tells us that E must be the same at all times, so dE/dt = 0.
We can use this information now to deduce the equation of motion. Differentiating the expression
for E with respect to time gives us

E = m x x +k xx = 0 . (16)
We can factor this equation into
x(m x +kx) = 0. (17)
so either x = 0, which tells us that the system may be at rest, or m x +kx = 0 which is the equation
of motion derived above. The situation in which the system never moves is not of great interest to
us in a course on dynamics. The other result is very useful, however, because it shows us how to
obtain an equation of motion for a system once we know the total energy of the system. Note that
we have assumed that the energy in the system is conserved. In some cases, friction and other forces
dissipate energy to the surrounding. In these cases, the energy of the system decreases in the course
of time and so the approach we have used here cannot be applied.
2.3 Free vibration of a SDF system with damping
In reality, many systems that we meet have damping. The damping can be caused by friction in
sliding parts or by viscous effects in a uid, for example. The mathematical specication of a
damping force can be quite tricky because of the different physical mechanisms that can cause
the damping. In all cases, however, the effects are similar: energy in the system is removed and
dissipated to the surroundings, often in the form of heat or noise. We will be primarily concerned
with viscous damping for which the mathematical specication is relatively simple. Just as with the
masses and springs, we will use lumped dampers to model the effects of damping in a system. This
means that although in reality the damping effects may be distributed throughout the system, in our
model the damping will occur at a well dened point.
2.3.1 Viscous damping and solution of the damped motion equation
The idea of viscous damping is that a system moving with a velocity v is slowed down by a force
proportional to v. Fast moving objects therefore encounter large forces, while slowly moving objects
10
encounter only small forces. The direction of the force is always to oppose the velocity so that the
system is always slowed by the damping, rather than speeded up. Mathematically, we can say that
the viscous damping force F is given by F =cv where c is a constant which determines the amount
of damping present in the system. Lets look at a simple system acting under the effects of viscous
damping.
x
k
m


c


Figure 7: A spring-mass-damper system.
Figure 7 shows how we represent symbolically a spring-mass system retarded by a viscous damper
(dashpot). Recalling that the force due to the viscous damper is cv =c x we can draw the free
body diagram for this system as shown in Figure 8.
x m

x c

x

m
kx
Figure 8: The free body diagram for the system of Figure 7.
11
Notice the direction of the forces. The positive x direction is from left to right in the gure, so if
the mass is accelerating from left to right both x and x are directed left to right also. As we have
already discussed, the inertial force acts to oppose the acceleration we attempt to give the mass, so
acts right to left. Also, the damping force acts to oppose the motion, and since the mass moves from
left to right for positive x, the damping force must act right to left. The equation of motion can now
be written down for this system. It is
m x +c x +kx = 0 . (18)
This is a linear, homogeneous second order equation. The solutions of this equation are of the form
exp(t). We can see why this is the case by thinking about the properties needed for a function to
be a solution of equation (18). The coefcients in the equation are constants, so if we can choose
a function that under differentiation is equal to itself multiplied by a constant, then this is a good
candidate for a solution to (18). The function that satises this need is of course the exponential
function. The constant must now be chosen so that equation (18) is satised. To this end, we
substitute x = exp(t), x = exp(t) and x =
2
exp(t) into (18). We get

2
+
c
m
+
k
m
= 0 . (19)
This quadratic equation (characteristic equation) has roots

=
c

c
2
4km
2m
, (20)
and the corresponding solutions are given by
x
+
(t) = exp(
+
t) ,
x

(t) = exp(

t) .
These are two complementary functions which solve equation (18). Since either of these functions
is a solution of (18), and the equation is linear, any combination Ax
+
+Bx

is also a solution (A and


B are arbitrary constants). The most general solution is therefore
x(t) = Aexp(
+
t) +Bexp(

t) . (21)
It is worth recognizing that the roots
+
and

and therefore also the complementary functions


x
+
(t) and x

(t) are not always real. Of course, the nal solution x(t) must be real, and this can be
ensured by taking, in a last step of the calculation, the real part of equation (21),
x(t) = Re[Aexp(
+
t) +Bexp(

t)] . (22)
The behaviour of this solution depends on whether the roots

of the characteristic equation are


real or complex.
12
If c
2
> 4km both roots given by equation (20) and, hence, also the complementary functions are
real. Inserting the values for
+
and

into equation (21) gives


x(t) = Aexp
__

c
2m
+
_
_
c
2m
_
2

k
m
_
t
_
+Bexp
__

c
2m

_
_
c
2m
_
2

k
m
_
t
_
. (23)
This expression can be re-arranged and x(t) can written in terms of hyperbolic functions:
x(t) = exp
_

c
2m
t
_
_
Ccosh
_
_
_
c
2m
_
2

k
m
t
_
+Dsinh
_
_
_
c
2m
_
2

k
m
t
__
(24)
where C = A+B and D = AB.
If on the other hand c
2
< 4km then the roots are complex and the values of exp(
+
t) and exp(

t)
will be complex quantities. The roots will be complex conjugates of each other, so the real part of
the roots is the same while the imaginary part is the same magnitude for each root, but of opposite
sign. We can therefore write

+
=
R
+i
I
,

=
R
i
I
where the real and imaginary parts
R
and
I
are

R
=
c
2m
,
I
=
_
k
m

_
c
2m
_
2
.
As for the real roots case, the general solution is given by equation (21), but now the exponentials
and the constants A, B must be envisaged as complex numbers, A = A
R
+iA
I
and B = B
R
+iB
I
.
Inserting in Eq. (21) gives
x(t) = Aexp(
+
t) +Bexp(

t)
= exp(
R
t)[(A
R
+iA
I
)exp(i
I
t) +(B
R
+iB
I
)exp(i
I
t)]
= exp(
R
t)[(A
R
+B
R
)cos(
I
t) (A
I
B
I
)sin(
I
t)
+i((A
I
+B
I
)cos(
I
t) +(A
R
B
R
)sin(
I
t))]
= exp(
c
2m
t)
_
E cos
_
_
k
m

_
c
2m
_
2
t
_
+F sin
_
_
k
m

_
c
2m
_
2
t
__
(25)
where in the last line I have taken the real part and used new constants E = A
R
+B
R
and F =
A
I
+B
I
to replace the old ones. Note the similarity to equation (24).
The form of the solutions (23) and (25) is very interesting. Equation (23) is the sum of two expo-
nential decays. This means that the motion of a system for which this equation is applicable, i.e.
when c
2
>4km, is just a decay towards zero. No oscillations can be seen, and we refer to the system
13
Figure 9: An example of the solution for an
over-damped system. The solution exhibits no
oscillations.
Figure 10: An example of the solution for an
under-damped system. The envelope of the os-
cillations is a decaying exponential.
as being over-damped. On the other hand, when c
2
< 4km, equation (25) shows that the motion
consists of an oscillatory part given by the sinusoidal functions in the brackets, multiplied by an ex-
ponential decay. The system therefore oscillates, but the amplitude of the oscillations decays. The
system is referred to as being under-damped. Figures 9 and 10 show plots of what these solutions
look like.
The two different types of behaviour, over-damped and under-damped oscillation, can be identied
using the ratio c
2
/4km. For c
2
/4km > 1 the system is over-damped and for c
2
/4km < 1 the system
is under-damped. For c
2
/4km = 1 the system is said to be critically damped. We wont look at this
case because in practice the ratio is never precisely equal to one.
2.3.2 Canonical form of the damped free vibration equation
We have seen how to solve the damped free vibration equation for a mass-spring damper system.
We considered the equation
m x +c x +kx = 0 . (26)
When we are dealing with other vibrating systems, the coefcients in this differential equation may
change, and even their physical meaning may be different. We will see an example of this in the
next section. However, the basic structure of the equation of motion and the method of solution
are always the same. Rather than nding the solution separately for each case, we write equations
of this type in a general form. This is usually referred to as canonical form. We can then use the
solution to the canonical equation straight away, rather than having to redo all the algebra. To this
end, we use two fundamental parameters. The rst parameter is the frequency of vibration of the
system without damping. We also refer to this as the natural frequency
0
of the system. For the
14
mass-spring system described by equation (26), we nd that the natural frequency is

0
=
_
k
m
. (27)
To characterize the damping, we introduce the damping ratio which determines whether the sys-
tem is under-damped or over-damped. For our mass-spring-damper system the damping ratio is
dened by

2
=
c
2
4km
. (28)
As we have seen in the preceding section, >1 gives over-damping, and <1 gives under-damping.
We can rearrange equation (28) to give a useful alternative expression for the damping ratio:
=
c
2

km
=
c
2m
0
. (29)
We can now write equation (26) in terms of and
0
instead of k, m and c. Dividing through by m
we have
x +
c
m
x +
k
m
= 0 . (30)
and from the relations for
0
and
k
m
=
2
0
,
c
m
= 2
0
, (31)
we nd that this can be written as
x +2
0
x +
2
0
x = 0 . (32)
By writing our equations in this canonical form, we can see at once the value of the damping
parameter and the undamped natural frequency. The true frequency of vibration is the coefcient of
t in the sine and cosine terms in equation (25). It is called the damped frequency
d
and is given by

d
=
_
k
m

_
c
2m
_
2
(33)
We can express this in terms of
0
and :

d
=
0
_
1
2
. (34)
This gives us the result that if damping is very small then
d

0
. We can also write the solution,
equation (25), using these new parameters:
x(t) = exp(
0
t)[Ecos(
d
t) +F sin(
d
t)] . (35)
Just as we did with the undamped system, we can rewrite the sum of the sine and cosine term as a
sine term with a phase shift. We get an equivalent formula to the above:
x(t) = Hexp(
0
t)sin(
d
t +) . (36)
15
The constants H and can be determined from the initial conditions.
Another parameter which is often used to characterize damped vibration is the logarithmic decre-
ment of the oscillation. The logarithmic decrement is dened as the natural logarithm of the ratio
of any two successive maxima of the oscillation. It is related to the damping ratio by
=
2

1
2
(37)
Exercise: 5
Use the solution (36) to derive this expression from the ratio between two succes-
sive maxima of the oscillating solution.
Despite all the mathematics above, we can see that the effect of damping on the vibrating system
is actually quite straightforward. If the damping is larger that a critical amount, given by = 1,
the system is over-damped and simply relaxes to its equilibrium position without any oscillation.
The more interesting situation from our point of view is when the damping is less than the critical
damping. In this case, the addition of the damping alters the frequency of the oscillations, and
makes their amplitude decay gradually to zero.
To use the canonical form to best advantage, we rst write down the equation of motion for a system
using the free body diagram. We can then identify the parameters
0
and , and from these deduce
whether the system is under or over-damped, the frequency of vibration and the rate of decay of the
oscillations (if present).
2.3.3 Rotational vibration
Until now we have considered systems where masses move along a given xed direction, i.e. the
motion is translational. However, in many important cases vibration is related to the rotation of
parts. In this case our procedure of setting up the equations of motion must be modied. Figure
11 shows a system which may exhibit rotational vibrations: A square block of mass m can rotate
around the axis A and is attached to the walls by a spring of constant k and a damper with damping
constant c. The spring constant is such that the block is at rest in the position shown.
Lets see what happens when we rotate the block to the right by a small angle . The top right corner
of the block (where the damper is attached) moves to the right by l sin and the bottom right corner
where the spring is attached moves downward by the same amount. This leads to a compression of
the spring and, if the rotation occurs at nite speed, to a viscous damping force from the damper.
To work out the spring force and the damping force, we use linearized relations by noting that, for
small angles , sin . Hence the spring and the damper are compressed by x = l. The velocity
of compression of the damper is obtained by taking the time derivative, x = l

.
The corresponding forces are drawn in Figure 12. Now we have to keep in mind that the block
16


c


k
m
l
l
B
A
Figure 11: A system undergoing rotational vibration.
rotates around the axis A, so we have to consider the sum of all moments with respect to this axis.
The spring and damper forces are kl and cl

, and their moments with respect to A are kl


2
and
cl
2

. In addition we have to consider the effect of inertia. This leads to a moment I


A
where I
A
is
the moment of inertia of the block with respect to the axis A and =

is the angular acceleration.
Exercise: 6
Calculate the moment of inertia of the block around the axis A. Assume that the
block is homogeneous. (Result: I
A
= 2ml
2
/3)
The sum of the moments around A must be zero. Hence we end up with the equation of motion
I
A

+cl
2

+kl
2
= 0 . (38)
or, after inserting I
A
= 2ml
2
/3 and dividing by 2l
2
/3
m

+(3c/2)

+(3k/2) = 0 . (39)
This equation is very similar to the equation of motion for the translational motion of a mass-spring-
17







A
B
m
B

cl
B kl







B
B
B
B

J
kl
cl
2
2
Figure 12: (Left) forces and (right) moments acting in the system of Fig. 11
damper system, equation (18), and can be written in the same canonical form. To this end, we divide
by the pre-factor of the term with the highest derivative to obtain

+
3c
2m

+
3k
2m
= 0 . (40)
We identify this with the general canonical form as given by equation (32) - here we have the angle
instead of x as the dependent variable, but everything else remains the same:

+2
0

+
2
0
= 0 . (41)
By comparing coefcients, we nd

2
0
=
3k
2m
, 2
0
=
3c
2m
, (42)
and, hence, the natural frequency and damping ratio for this system are given by

0
=
_
3k
2m
, =
_
3c
2
8km
. (43)
We can now use the general solution of the canonical vibration equation by inserting these param-
eters into equation (35) or (36) and determining the remaining unknown parameters from initial
conditions.
18
x
c
m f(t)
k
Figure 13: An example of a simple system subjected to forcing
2.4 Periodically forced vibration
We now consider the vibration of a system which involves an external force. We rst consider
periodic forcing, where we will make use of the complex exponential method once again. You may
have found it possible to work through the problems so far without using complex exponentials. You
will struggle to do this for forced vibration. Please make sure you are familiar with the workings of
the complex exponential method, and make sure you can do the questions Ive set on it. Once you
are happy with the method, youll probably nd that the material that follows is not too complicated.
2.4.1 Steady-state response of a periodically forced system
Consider the system shown in Figure 13. We assume that the forcing f (t) is a periodic function, so
f (t) = f
0
cos(t). The equation of motion for the system shown in Figure 13 is
m x +c x +kx = f (t) = f
0
cos(t) . (44)
Before we solve this equation we consider the simplest case where the applied force is a constant
( = 0). A constant force leads to a constant elongation of the spring by x
0
= f
0
/k. We call x
0
the
static response of the system.
Now we put the equation into canonical form. We rst divide by m to get
x +2
0
x +
2
0
x =
f
0
m
cos(t) . (45)
The right-hand side of this equation can be written in terms of the static response and the natural
frequency:
x +2
0
x +
2
0
x =
2
0
x
0
cos(t) . (46)
19
This is the form which we will always use in the following. Looking at the problem from a math-
ematical point of view, I can see that the equation of motion is an inhomogeneous equation and so
the solution will be given by the sum of two complementary functions and a particular integral. The
complementary functions are the solutions of the homogeneous problem and have been discussed in
the previous section, they will both be damped oscillations, so after a reasonable time they will have
decayed to small values. At this stage, Im interested in how the system responds after quite some
time so that these initial effects will have died away. Thats what we call the steady-state response
of the system. I will therefore ignore the complementary functions in what follows.
Now let us take a closer look at equation (46). The left-hand side (LHS) basically involves x mul-
tiplied by various constants and differentiated. The solutions to this sort of equation are things like
x exp(t), where may be complex. The right-hand side (RHS) is a cosine, but I can write that
also in complex exponential form
x +2
0
x +
2
0
x =
2
0
x
0
exp(it) (47)
and regard x as a complex variable. I can recover the original equation of motion by looking at just
the real part of this equation. Since the equation is linear, I can say that if x = x
R
+ix
I
is a solution
then x
R
is a solution to the real part of the equation and x
I
is a solution to the imaginary part. I can
therefore go ahead and solve for the complex variable x and take the real part of x at the end of the
calculation. This real part will be a solution to the original equation of motion.
I expect the steady-state response to be
x(t) = X exp(it) , (48)
where x and X are complex. Substituting this expression gives
X =
x
0

2
0

2
+2i
0
+
2
0
. (49)
To bring this complex amplitude into a more useful form, I use that any complex number A+iB can
also be written in the form of a complex exponential:
A+iB = r exp(i) (50)
where r
2
= A
2
+B
2
and tan = B/A. When we write X in equation (49) in this form, we get
X =
x
0
_
_
1
2
/
2
0
_
2
+4
2

2
/
2
0
exp(i) , (51)
where
=arctan
_
2/
0
1
2
/
2
0
_
. (52)
20
We now insert insert X into the solution (48):
x(t) =
x
0
_
_
1
2
/
2
0
_
2
+4
2

2
/
2
0
exp(i[t +]) , (53)
As I explained above, this is the solution to the complex form of the equation of motion. The
physical solution is simply the real part of it. I use that
cos(t) = Re[exp(it)] . (54)
and nd
x
R
=
x
0
_
_
1
2
/
2
0

2
+4
2

2
/
2
0
cos(t +) . (55)
What does this mean? There are several interesting points to extract from this solution. Firstly, it is
basically another sinusoidal function of the form
x = Acos(t +) , (56)
where the amplitude of the vibration is
A =
x
0
_
_
1
2
/
2
0
_
2
+4
2

2
/
2
0
(57)
and the frequency is the same as that of the forcing. However, there is a phase shift between the
forcing and the response. The system therefore responds to the forcing by vibrating at the same
frequency but with a delay,
=arctan
_
2/
0
1
2
/
2
0
_
. (58)
The phase lag depends on the frequency of forcing. For small , is small. For =
0
,
=/2, and for , .
Exercise: 8
What does a phase shift of radians mean physically?
Exercise: 9
Write down the values of for = 0.1 and = 0,
0
+,
0
, where is a
very small number. Verify that the values of the phase shift for small are 0,
for =
0
, =/2, and for , .
When you use your pocket calculator to get the phase shift , you must be careful when taking the
inverse tangent of a negative quantity. The inverse tangent is only dened up to a factor of n where
n is an integer number, and you may have to subtract to get the phase shift right.
21
We now look how the amplitude behaves as a function of . Take the expression for A,
A =
x
0
_
_
1
2
/
2
0
_
2
+4
2

2
/
2
0
(59)
and consider the case when the damping ratio is small. Now, the squared terms inside the square
root are always positive, so since they appear in the denominator, the amplitude is largest when they
are small. The rst term is zero when =
0
. So when the damping is small and the forcing is
the same frequency as the frequency of free vibration, the amplitude is a maximum. This is called
resonance. For small values of the damping ratio we can often use the approximate expression
A =
x
0
|1
2
/
2
0
|
. (60)
This expression shows that the resonance peak is located at approximately
0
(this follows from
the small damping ratio assumed in equation (60)). The amplitude of the dynamic response at
resonance for small is 1/(2).
Exercise: 10
Calculate the frequency and amplitude of resonance (the frequency and amplitude
where the response has its maximum) for the case where is not small.
Most of what weve been dealing with here can be neatly expressed in two graphs; one for the am-
plitude and one for the phase as functions of the frequency of forcing. The graph for the amplitude
is probably the most useful form of the frequency-response graph. It tells us a great deal about the
dynamics of a system, all on a single plot. If we plot on (x, y) axes, then we let
y = A/x
0
, x = /
0
, (61)
and so from equation (59) we have
y =
1
_
(1x
2
)
2
+4
2
x
2
(62)
which determines the shape of the frequency-response graph. Because at resonance the response is
usually very large, you will often nd these graphs plotted in logarithmic coordinates (i.e. plot lny
against lnx).
Exercise: 11
Use MATLAB or Mathcad to deduce the shape of these graphs for a number of
values of and add sketches to your notes. Deduce the height and width of the
resonance peak and label it.
2.4.2 Force transmission and vibration damping
We now consider the following question: Given the amplitude f
0
of a periodic forcing acting on the
mass m in Figure 13, what is the amplitude of the force transmitted to the support?
22
Adding the spring and damper forces, we see that the transmitted force is given by
f
s
(t) = kx(t) +c x(t) (63)
Inserting the solution given by equation (56), we nd
f
s
(t) = kAcos(t +) cAsin(t +) (64)
This is a periodic function, and at the moment we are interested only in the amplitude. If we shift
the time axis according to t

=t / we nd that
f
s
(t

) = f
s,0
sin(t

) (65)
where tan

=k/(c) and f
s,0
=

k
2
A
2
+c
2

2
A
2
. Hence the amplitude of the transmitted force
is
f
s,0
= A
_
k
2
+
2
c
2
=
x
0

k
2
+
2
c
2
_
_
1
2
/
2
0
_
2
+4
2

2
/
2
0
(66)
and using the relations x
0
= f
0
/k, c
2
= 4km
2
and
2
0
= k/m this can be written as
f
s,0
= f
0

_
1+4
2

2
/
2
0
_
1
2
/
2
0
_
2
+4
2

2
/
2
0
= f
0
TR (67)
The ratio between the force acting on the mass and the force transmitted to the surroundings is
called the transmissibility ratio
TR =

_
1+4
2

2
/
2
0
_
1
2
/
2
0
_
2
+4
2

2
/
2
0
. (68)
Exercise: 12
Sketch the curves of TR plotted against x =/
0
for x = 0. . . 5 in intervals of 0.5.
Use values of = 0.1, 0.2, 0.5 and compare the different dampings. Plot the same
graph on logarithmic scales.
The form of the transmissibility ratio shows that at the resonant frequency, large values of damping
are good because they reduce force transmission. However, at high frequencies the opposite is the
case and the most efcient way to reduce force transmission is to make
0
as small as possible (soft
springs, large masses). In practical circumstances it is vital to be aware of this compromise.
2.4.3 Moving boundary condition and amplitude transmission
Until now we have considered forced vibration where an external force is acting directly on a vi-
brating mass. A different type of excitation is shown in Figure 14.
23
x
c
m
k
v(t)
Figure 14: A system with moving boundary excitation
We assume that the support of the system moves periodically according to y(t) =y
0
cos(t) and ask
for the amplitude of the steady-state response of the system.
The equation of motion is
m x +c( x y) +k(x y) = 0. (69)
We collect the terms involving zy on the right-hand side and bring the equation into canonical form
by dividing through m
x +2
0
x +
2
0
x =2
0
y
0
sin(t) +
2
0
y
0
cos(t). (70)
We now bring also the right-hand side into canonical form. To this end, we write
2
0
y
0
sin(t) +
2
0
y
0
cos(t)
=
2
0
y
0
[cos(t) (2/
0
)sin(t)]
=
2
0
[y
0
_
1+4
2

2
/
2
0
] sin(t +) . (71)
We nally shift the time axis to get rid of the phase and nd that the equation of motion assumes
the canonical form
x +2
0
x +
2
0
x =
2
0
x
0
cos(t

) (72)
with x
0
= y
0
_
1+4
2

2
/
2
0
. We can now use the solution (56), x(t) = Acos(t +) with the
amplitude A given by (59). As a result we nd that the ratio of the vibration amplitudes of the mass
and of the support is given by
A/y
0
= TR (73)
with the same transmissibility TR as in the previous section. Hence, equation (68) gives us both
force and amplitude transmission, and what has been said about the inuence of the parameters
0
and on vibration isolation also applies to the present case.
24
2.5 Applications
2.5.1 Accelerometer
As an application of the theory of forced vibration with moving boundary excitation, we consider a
device known as an accelerometer. As the name suggests, it is used to measure the acceleration of
a vibrating surface. Essentially, the device consists simply of a seismic mass, basically a lump of
metal, that is attached by a spring damper system to the surface whose vibration is to be measured.
A sensor is able to measure the distance between the seismic mass and the surface. Our task is to
relate the extension of the spring, which we can measure, to the acceleration of the surface. Figure
15 shows a diagram of the situation.
x
c
m
k
v(t)
Figure 15: A schematic diagram of an accelerometer. x measures distance from a xed point in
space to the mass, y measures distance from a xed point in space to the vibrating surface.
The equation of motion is
m x +c( x y) +k(x y) = 0 . (74)
We now introduce the new variable z = x y. We can differentiate to nd
z = x y , (75)
z = x y , (76)
and then eliminate x from the equation of motion in favour of z. We get
m z +c z +kz =m y , (77)
which looks quite familiar. Let us assume that the surface is vibrating with frequency so that
y(t) =Y cos(t) (78)
25
and so
y =
2
Y cos(t) , (79)
which we can use in our equation of motion to give
m z +c z +kz = m
2
Y cos(t) , (80)
which is in the form which we are used to solving, with the amplitude of the forcing given by m
2
Y.
Dividing through by m we have
z +2
0
z +
2
0
z =
2
Y cos(t) , (81)
We know how to solve this sort of equation. The canonical form is
z +2
0
z +
2
0
z =
2
z
0
cos(t) , (82)
By comparison, we nd that the static response z
0
is
z
0
=Y
2
/
2
0
(83)
and the amplitude of the dynamic response is
Z =
z
0
_
(1(/
0
)
2
)
2
+4
2
(/
0
)
2
=
Y(/
0
)
2
_
(1(/
0
)
2
)
2
+4
2
(/
0
)
2
. (84)
Hence, the ratio of the amplitudes of vibration of the accelerometer and the surface is
Z
Y
=
(/
0
)
2
_
(1(/
0
)
2
)
2
+4
2
(/
0
)
2
. (85)
If we make the free vibration frequency
0
very high by a suitable choice of k and m, then /
0
will be small and
Z
Y
=

2

2
0
(86)
to a good approximation. Therefore we can write

2
Y =
2
0
Z . (87)
Now, the surface is moving sinusoidally according to
y =Y cos(t) (88)
and so the acceleration of the surface is
y =
2
Y cos(t) (89)
and the amplitude of this acceleration is
A =
2
Y (90)
26
Therefore,
A =
2
0
Z . (91)
Since we can measure the amplitude Z and we already know the frequency
0
, we can measure the
acceleration amplitude A. Since we are far below the free vibration frequency, our measured signal
z(t) will be in phase with the term on the right-hand side of Eq. (81) and therefore in antiphase with
the acceleration.
A further degree of sophistication is to add damping to the accelerometer in order to improve its
accuracy. The whole idea in going from Eq. (85) to (86) is that the term under the square root
should be as close as possible to 1. This is achieved by making /
0
as small as possible. To do
even better, we retain an extra term of order
2
/
2
0
:
Z
Y
=
(/
0
)
2
_
(1(/
0
)
2
)
2
+4
2
(/
0
)
2

2
0
[1(2
2
1)(/
2
0
)] . (92)
This time we have kept extra terms to which is why our original result (86) is not exact. These terms
become signicant when
2
/
2
0
is not so small. However, we can get rid of them! If we set =
1/

2 then the above equation reduces to equation (86). We can therefore expect our accelerometer
to be more accurate if we apply damping so that 0.7.
Many accelerometers work on this principle. Low frequency accelerometers use a damping ratio
of 0.7 as described above. This also improves the phase distortion; you can read a description of
phase distortion in Thompsons book if youre interested. The ones you will meet in the lab use a
piezoelectric crystal which has a very high natural frequency and therefore there is no need for any
damping (can you explain why?). These accelerometers are more suited to high frequency work.
2.5.2 Rotating out of balance rotor
Many machines have rotating parts which can vibrate. The system we will look at are the vibrations
of a machine which are caused by a rotor which is out of balance. Figure 16 shows a machine of
mass M which contains a rotor of mass m with an eccentricity of e. The machine is mounted on
a solid oor by a spring of stiffness k and a damper with damping constant c and can move in the
vertical (x) direction only.
From the free-body diagram shown in Figure 17, we nd that the equation of motion for the machine
and rotor in the x direction is
(m+M) x +c x +kx = me
2
cos(t) . (93)
27
Mt
G
e O
x(t)
m M
c k
Figure 16: Schematic diagram of a machine with an out of balance rotor. The centre of mass of the
rotor is located at G.
We follow our usual procedure of casting this into canonical form,
x +2
0
x +
2
0
x =
2
0
x
0
cos(t) . (94)
where now

0
=
_
k
m+M
, =

c
2
4k(m+M)
, x
0
=
me
m+M

2
0
. (95)
It follows fromthe results of Section 2.4.1 that the steady-state solution is given by x(t) =X cos(t +
) where the amplitude X is given by
X =
me
m+M

2
/
2
0
_
(1
2
/
2
0
)
2
+4
2

2
/
2
0
(96)
and the phase is given by
tan =
2/
0
1
2
/
2
0
. (97)
If we look at the amplitude and the phase of the motion for low, resonant and high frequencies we
see that

0
X 0 0 ,
=
0
X = me/[2(m+M)] =/2 ,

0
X me/(m+M) . (98)
28
me

m x(t)

c M kx x x
2
M
x
Figure 17: Free body diagram for the system of Figure 16.
The highest possible amplitude occurs at resonance in a system for which m M. In this case, the
maximum amplitude is X = e/. If the damping is small this can be very large.
Exercise: 13
Use the results of the previous sections to determine the amplitude of the force
which is transmitted to the ground. What is the force transmitted at resonance?
2.5.3 Out of balance in reciprocating internal combustion engines
The inertial forces due to the motion of the piston and con rod in an engine must be balanced by
forces acting on the engine and these can cause vibration. In this section we will deduce what these
forces are and what measures can be taken to keep them to a minimum. Figure 18 shows a schematic
diagram of a piston in a cylinder a distance x from bottom dead centre attached by a con rod AB to
a crank shaft BC which rotates about C. The length of the shaft is r and the length of the con rod is
A | u C
B
l r
x lr-x
Figure 18: Schematic diagram of a piston, con rod AB and crank BC.
29
l. The distance between the piston and the crank axis is therefore r +l x as shown.
To nd the inertial force due to the motion of the piston, con rod and crank shaft, we will seek to
relate x to , the angle through which the crank shaft has rotated. If the engine runs at a constant
speed then

, the angular velocity of the crankshaft, is a constant .
The problem is that because of the geometry the motion x(t) of the piston is not simply a sinusoidal
function. To work out what it is, we project the lengths of AB and BC onto AC to give
ABcos+BCcos = AC (99)
and putting in the various terms gives us
l cos+r cos = l +r x (100)
From the sine law, we have
l sin = r sin (101)
therefore we can eliminate because
cos =
_
1sin
2
=

1
r
2
sin
2

l
2
. (102)
hence
x = r +l r cosl

1
r
2
sin
2

l
2
. (103)
If we introduce a parameter = r/l then we have
x = r
_
1+
1

cos
1

_
1
2
sin
2

_
. (104)
In an engine, the value of is usually around 1/3 to 1/4 and so we can expand the square root as a
power series and ignore higher powers of because they will be small. For small , we have from
the Taylor series that
_
1
2
sin
2
1

2
sin
2

2
, (105)
so we can write
x = r
_
1cos+
sin
2

2
_
. (106)
where I have ignored powers of higher than one. This approximation is usually good enough. We
now want to differentiate twice with respect to t in order to nd the acceleration. We assume that
the engine is running at constant speed, = t. We insert this into x(t) and use that
sin(t)cos(t) =
1
2
sin(2t) . (107)
30
Taking the derivative of x(t) gives
x = r[sin(t) +sin(t)cos(t)] = r[sin(t) +

2
sin(2t)] , (108)
and
x = r
2
[cos(t) +cos(2t)] . (109)
Now we have calculated the acceleration of the piston, let us consider the forces acting in the system
of Figure 18. To make the analysis easier, we will model the con rod in an approximate way. The
true con rod shape is rather complicated, and we will replace it in our lumped model with two
masses, m
1
and m
2
, one at each end of the real rod. Figure 19 shows the con rod and its model.
Figure 19: Schematic diagram of the con rod and its model. G is the centre of gravity.
We choose the lengths a and b and the masses m
1
and m
2
so that m
1
+m
2
= M
CR
, where M
CR
is the
mass of the con rod, and m
1
a = m
2
b so the the centre of gravity of the con rod and the model of the
con rod are the same. Figure 20 shows the forces acting on the system with the model for the con
rod inserted. In order to balance the system, the size and shape of the crank is chosen so the rotation
of the mass labelled m
3
causes the same centrifugal force as the rotation of the mass m
2
from the
con rod. These forces balance, but the remaining inertial force M
1
x along the axis of motion of the
piston, where M
1
= M
P
+m
1
and M
P
is the mass of the piston, causes a force F on the crankshaft
and hence on the engine.
To summarise this section, we have looked at the motion of the piston, con rod and crankshaft
system at a constant angular velocity and deduced the inertial force due to the motion of the piston
and the con rod. The crankshaft shape is chosen to eliminate the non-axial component of the con
rod inertial force and the resulting force on the engine is given by
F = (M
P
+m
1
) x = (M
1
)r
2
[cos(t) +cos(2t)] (110)
31
M
1
x
m
2
rO
2
m
2
rO
2
m
3
M
p
m
1

Figure 20: Forces acting on the piston, con rod and crank.
The term M
1
r
2
cos(t) is called the primary unbalance force and the term M
1
r
2
cos(2t) is
the secondary unbalance force. To understand the origin of the secondary unbalance force, it is
useful to have a look at Figure 21. The motion x() of the piston as calculated from Eq. (104) is not
strictly sinusoidal, and the difference with a pure cosine can be well approximated by a cosine with
twice the frequency. The corresponding acceleration gives rise to the secondary unbalance force.
0 2 4 6 8 10
0.0
0.5
1.0
1.5
2.0

x
/
r
Figure 21: Upper full line: motion of the piston as a function of the angle of rotation of the
crankshaft calculated from Eq. (104) for = 1/4. Dotted line: approximation by a cosine. Lower
full line: Difference between the actual motion and a cosine.
32
We nowlook at the assembly of cylinders in an engine, and the resultant forces which act on the body
of the engine. The goal will be to devise an engine that is balanced, so the unbalance forces from
each cylinder cancel out. We will consider a four cylinder inline arrangement. Other arrangements
can be analysed by the same method.
Figure 22: Four cylinder inline arrangement. The cranks are attached at angles of 0, 180, 180 and 0
degrees.
Figure 22 illustrates a four cylinder engine where the pistons are connected to the crankshaft at
angles of 0, , and 0 radians (1-3-4-2 ring sequence). We want to know what the forces and
moments acting on the engine are. We will consider a general case, then see how this applies to the
four cylinder engine.
Suppose that there are N cylinders in a line, and that the i-th crank is attached at an angle of
i
.
After time t, the shaft rotates through an angle t and the crank is at an angle t +
i
to the vertical.
The force due to the piston and con rod is
F
i
= M
1
r
2
[cos(t +
i
) +cos2(t +
i
)] (111)
The total force is the sum over all the cylinders, so
F =

i
F
i
(112)
and is composed of a primary unbalance force oscillating at angular frequency , and a secondary
force oscillating at 2. By adding the contributions from all cylinders, we nd that the total primary
33
unbalance force F
P
is given by
F
P
= M
1
r
2

i
cos(t +
i
) = M
1
r
2
[cost

i
cos
i
sint

i
sin
i
] (113)
and the total secondary unbalance force is
F
S
= M
1
r
2

i
cos2(t +
i
) = M
1
r
2
[cos2t

i
cos2
i
sin2t

i
sin2
i
] . (114)
We also want to know the moments acting on the engine. If we take moments about a point O (the
center of gravity of the engine), and each crank i is a vertical distance z
i
from O, the primary and
secondary moments are
M
P
= M
1
r
2

i
z
i
cos(t +
i
) = M
1
r
2
[cost

i
z
i
cos
i
sint

i
z
i
sin
i
] (115)
and
M
S
= M
1
r
2

i
z
i
cos2(t +
i
) = M
1
r
2
[cos2t

i
z
i
cos2
i
sin2t

i
z
i
sin2
i
] . (116)
We can now evaluate the primary and secondary moments and forces from these expressions for
any inline arrangement of cylinders. For the four cylinder example, we have i = 1, 2, 3, 4 and
z
1
b, z
2
= a, z
3
= a and z
4
= b. We can evaluate the sums of the various terms as shown in
Table 1. The terms in the columns are the terms in the sums in the equations above.
crank
i
z
i
sin
i
cos
i
sin2
i
cos2
i
1 0 -b 0 1 0 1
2 -a 0 -1 0 1
3 a 0 -1 0 1
4 0 b 0 1 0 1
total 0 0 0 4
crank
i
z
i
z
i
sin
i
z
i
cos
i
z
i
sin2
i
z
i
cos2
i
1 0 -b 0 -b 0 -b
2 -a 0 a 0 -a
3 a 0 -a 0 a
4 0 b 0 b 0 b
total 0 0 0 0
Table 1: Sums for the primary and secondary unbalance forces and moments in a four cylinder
inline engine as shown in Figure 22
34
Using the table we can see that the primary and secondary unbalance momentsw as well as the
primary unbalance forces cancel, but the secondary unbalance force remains. The amplitude of this
is given by
F
0
S
= 4M
1
r
2
. (117)
This force can in principle be countered using out-of-balance rotating shafts. This adds to the
complexity of the engine, but ensures that it does not vibrate excessively.
Exercise: 14
Another possible solution to the problem of minimising vibration of a 4-cylinder
inline engine could be to use angles of 0
o
, 90
o
, 270
o
, 180
o
. Draw your own version
of Table 1 and conrm that in this case the primary and secondary unbalance forces
and primary unbalance moments cancel. Why is this design not commonly used?
Exercise: 15
Demonstrate that for 6 inline cylinders attached at angles of
0
o
, 120
o
, 240
o
, 240
o
, 120
o
and 0
o
and at distances of 2.5, 1.5, 0.5, 0.5, 1.5, 2.5
from the centre of the shaft, the primary and secondary unbalance forces and
moments all cancel.
35
2.6 Non-periodic forcing
In certain cases we wish to analyse the response of a system to non-periodic forcing. One example
of this might be a sudden shock. We will consider a shock rst of all, and then show that more
complicated forcings can be considered as a amalgamation of shocks one after the other. This will
give us a formula with which we can calculate the response to an arbitrary forcing.
Force



~ 1 /c

c Time
Figure 23: A short-lasting force.
Consider a spring-damper system such as that shown above in Figure 13, with the forcing function
shown in Figure 23. The duration of the force shown is given by which we will take to be small.
This means that the force lasts only briey, but is large ( 1/). Such a shock force can be envisaged
as a heavy kick. The strength of the kick can be characterized by its impulse which is dened as the
integral
I =
_

0
f (t)dt . (118)
What is the response of the system to an impulsive shock like this? Lets be clear about the problem
we wish to solve. Given the spring-damper system of Figure 13 subjected to an impulse of size I,
what is its response? We will assume that initially the system is in its equilibrium state.
Its actually quite straight-forward to work out. We will nd that the effect of the forcing is just to
give the system an initial kick, and from then on it will behave like a freely vibrating system. If we
can work out the size of this kick, this will allow us to set the initial condition for the subsequent
freely vibrating behaviour.
36
All we need to do is to recognize that Newtons second law can be written as
F
m
=
dp
dt
(119)
where p =mv is the momentum, and F
m
is the force acting on the mass m. Therefore, the momentum
of the mass is given by
p(t) p(0) =
_
t
0
F
m
dt . (120)
When the initial kick is very strong and very short, the effect of the springs and dampers during the
kick can be ignored because the external force is much larger than the spring and damper forces.
Initially, v(0) = 0 because the system starts at rest, so p(0) = 0 and the momentum of the mass m
just after the shock is given by
p(t = ) = mv() = I , (121)
and so
v(t = ) = I/m . (122)
Now, if is very small (the kick is very short), we can consider the motion of the system to start
at time 0 with the velocity v = I/m, and since the forcing is zero after time we need only consider
the motion of the freely vibrating system. We found above that the solution for a freely vibrating
system is
x(t) = exp(
0
t)[Acos(
d
t) +Bsin(
d
t)] (123)
and if we set the initial condition x(0) = 0, it follows that A = 0, i.e.,
x(t) = Bexp(
0
t)sin(
d
t) . (124)
The other initial condition is x(0) = I/m. We nd:
x(t) = Bexp(
0
t)[
0
sin(
d
t) +
d
cos(
d
t)] , (125)
so
x(0) = B
d
= I/m , (126)
hence
B =
I
m
d
, (127)
and the solution we require is
x(t) =
I
m
d
exp(
0
t)sin(
d
t) . (128)
This expression gives us the response of a system to an impulse of size I delivered at time t = 0. It
is usual to dene the unit impulse response function h(t) to be
h(t) =
1
m
d
exp(
0
t)sin(
d
t) , (129)
37
and so the response to an impulse I is
x(t) = Ih(t) . (130)
We often nd that we can approximate the behaviour of other forces as shocks, even if they last for
quite a long time. The spring-mass system has a time scale for its vibrations of T
0
= 2/
0
and if
the time over which a force acts is much smaller than T
0
we can usually treat it as a shock and write
down the response using the unit impulse response function h(t).
For complicated non-periodic forcing, we can generalise the above method. Consider the forcing
function shown in Figure 24:
Force




: t-:


: :d: t Time
Figure 24: A general non-periodic forcing.
We can imagine that the forcing F() is split up into lots of pieces dened over intervals of time d
(note that for clarity Im using the variable to denote time rather than t - the reason will become
clear later). Each of these can be thought of as a little impulsive shock. If we consider the response
at time t, we can think of this response as being built up of many responses to the previous shocks
- remember, we are allowed to build up solutions like this because the equation is linear and so the
super-position principle holds.
Lets consider the response at time t due to the shock shown at time . The impulse due to this shock
is
I

= F()d (131)
Now, the effect of this shock at time t depends on the time t that is elapsed after the shock. It is
given by
x

= h(t )I

= h(t )F()d , (132)


where Ive used the subscript

to denote that this is the part of the response due to the shock at time
38
. The full response is the total of all the shocks from time 0 to time t, i.e.,
x(t) =
_
t
0
x

d =
_
t
0
F()h(t )d , (133)
and so we can calculate the response using this expression, provided we can do the integral. In full,
the formula for the response is
x(t) =
_
t
0
F()
m
d
exp(
0
(t ))sin(
d
(t ))d (134)
which is called the convolution formula. For the case of small or no damping, this expression
simplies a bit to
x(t) =
_
t
0
F()
m
0
sin(
0
(t ))d (135)
which is in general easier to integrate. This I will refer to as the convolution formula without
damping.
2.6.1 Shock damping and shock response spectrum
To assess whether a vibrating system is capable of effectively damping shocks, we consider the
shock response spectrum. This is a plot relating the response to the forcing duration, but plotted
in a clever choice of variables: On the y-axis, the ratio of the peak response x and the peak static
response x
0
is plotted. The peak response is the highest maximum of the response to the shock,
which is also sometimes called the maximax. The peak static response is just the response that
would be caused by a force equal to the peak value of the forcing applied statically. On the x-axis,
the frequency of free vibration divided by the frequency scale of the forcing is plotted. For a force
of duration T
s
, the frequency scale is simply 2/T
s
. Therefore, if the peak value of the shock force
is F
0
and the system has a mass m, the peak static response is
F
0
/k = F
0
/(m
2
0
) . (136)
If the shock force lasts for a time T
s
then the frequency of free vibration of the system divided by
the frequency scale of the forcing is
0
T
s
/(2). On the shock response spectrum, we plot xm
2
0
/F
0
against
0
T
s
/(2) .
The shock response spectrum depends in general on the shape of the force pulse. As an example,
we consider the response of an undamped mass-spring system to a rectangular pulse,
F() =
_
F
0
, 0 T
s
0 else .
(137)
The response of an undamped spring-mass system to this pulse is obtained from equation (135). We
consider separately the two cases t T
s
and t > T
s
:
39
0.0 0.5 1.0 1.5 2.0
0
1
2


P
e
a
k

a
m
p
l
i
t
u
d
e


[
F
0
/
(
m

0
2
)
]
Shock duration [2/
0
]
Figure 25: Shock response spectrum for a rectangular pulse. Full line: Primary SRS; Dashed line:
Residual SRS
(i) For t T
s
the response is
x(t) =
F
0
m
0
_
t
0
sin(
0
(t ))d =
F
0
m
2
0
[1cos(
0
t)] . (138)
The maximum of the response is at t
max
= (2n
max
1)/
0
with the peak value x = 2F
0
/m
2
0
. n
max
is a positive integer number. Hence, this maximum is reached only if T
s
> /
0
.
(ii) For t > T
s
we nd
x(t) =
F
0
m
0
_
T
s
0
sin(
0
(t ))d =
F
0
m
2
0
[cos(
0
(t T
s
)) cos(
0
t)] . (139)
The maxima are in this case given by
x =
2F
0
m
2
0
| sin(
0
T
s
/2)| . (140)
The shock response spectrum (SRS) is plotted in Figure 25. The primary SRS shows the absolute
maximum of the response which occurs during the time of the shock, and the residual SRS shows
the absolute maximum of the response occurring after the end of the shock. It is interesting to note
that for shock durations which are integer multiples of the period of vibration, there is no residual
response at all.
40
We now determine the conditions under which the maximum of the shock response falls below the
static response, i.e. where we have shock damping. This requires that
x = sin(
0
T
s
/2) < 1/2 . (141)
Hence, shock damping is achieved if the pulse is sufciently short such that
0
T
s
< /3, i.e. the
ratio between the shock duration and the period of free vibration must be T
s
/T < 1/6. We may
compare this with the criterion for vibration damping which follows from equation (67), where we
have observed that vibration damping is achieved if
0
/ <

2, i.e., the ratio between the period


of the forcing and the period of free vibration must be less than 1/

2. In either case an efcient


strategy to achieve damping is to make the natural frequency of the system as small as possible
(large masses and/or soft springs).
41

x
1
x
2
k k k
m 2m
Figure 26: A two degree of freedom system consisting of two masses connected by springs.
3 The theory of systems with many degrees of freedom
The generalisation of what we have done for one degree of freedom systems to many degrees of
freedom is straight-forward provided we do it in the right way. What we will nd is that if we use
matrices to represent our equations, then the free body diagram equation of motion complex
exponential method we have used works just as well for the many degree of freedom systems. The
main complication is the maths, which can get rather lengthy. In fact, we will often have to resort
to computers and numerical methods in order to solve the equations. We will also nd there are one
or two extra things going on in these systems that simply do not occur in a single degree of freedom
system.
3.1 Free vibration
We will start our study of multiple degree of freedom systems with a relatively simple example
shown in Figure 26.
Exercise: 16
We have discussed in the Introduction how to obtain equations of motion for this
kind of system. Draw a free body diagram and write down the equation of motion
for each mass.
If we write down the equations of motion for the two masses in the system shown and collect similar
terms together we get a set of two coupled linear differential equations:
42
m x
1
+2kx
1
kx
2
= 0 ,
2m x
2
kx
1
+2kx
2
= 0 . (142)
I can represent these equations of motion in matrix form:
_
m 0
0 2m
__
x
1
x
2
_
+
_
2k k
k 2k
__
x
1
x
2
_
= 0 . (143)
Exercise: 17
Multiply out the matrix equation above to check that you do recover the two equa-
tions of motion.
I could write this matrix equation symbolically as
M

x +Kx = 0 , (144)
where bold quantities represent matrices and the arrows represent vectors. M and K are the mass
and the stiffness matrices. They are given by
M =
_
m 0
0 2m
_
,
K=
_
2k k
k 2k
_
. (145)
Written in this form, you can see that the equation looks really similar to the single degree of
freedom equations we are used to solving. I can solve the equation using the complex exponential
method. I write
x =

Aexp(it) , (146)
where the usual x and A quantities I get in the single degree of freedom systems have become
vectors. Alternatively, I can write out this equation in long hand as the pair
x
1
= A
1
exp(it) ,
x
2
= A
2
exp(it) . (147)
I calculate the derivatives and substitute into the equation of motion in the usual way to get

2
M

Aexp(it) +K

Aexp(it) = 0 . (148)
Cancelling out the exponentials and collecting terms gives me
(
2
M+K)

A = 0 , (149)
43
which I can write as
B

A = 0 (150)
with
B =
2
M+K . (151)
It is important to understand that this is just a set of two coupled linear equations for the unknown
components of the vector

A, i.e., for the amplitudes of vibration of the two masses. Now a general
result in the theory of linear algebra is that for the system of equations B

A = 0 to have nontrivial
solutions, it must be true that the determinant of B must be zero,
|B| = 0 . (152)
If this does not ring a bell, please take the time to go over your maths notes to refresh you memory.
Replacing B, we have deduced at this point in the calculation that
|
2
M+K| = 0 (153)
If I now put in the expressions for the matrices I get the determinant

2k m
2
k
k 2k 2m
2

= 0 . (154)
Now I can expand this determinant out to get
(2k m
2
)(2k 2m
2
) k
2
= 0 (155)
which rearranges to
2m
2

4
6km
2
+3k
2
= 0 , (156)
which is a quadratic equation in
2
. I can solve the quadratic equation to nd two roots. I get

=
k
m
_
3
2

3
2
_
. (157)
Negative frequencies dont really mean anything. I therefore have two solutions which I will label

+
and

. They are given by

+
=

_
k
m
_
3
2
+

3
2
_
, (158)

_
k
m
_
3
2

3
2
_
. (159)
This result is important. I have managed to calculate the frequencies of free vibration of the system
of two masses. Interestingly, there are two possible frequencies for free vibration. Either can occur
44
i.e. the system will resonate at either of these frequencies. More generally, a system with n degrees
of freedom has n frequencies of free vibration.
Now I have evaluated , I can go back to equation (149) and put in either
+
or

. For each value


I can use the matrix equation to calculate a ratio A
1
/A
2
.
Exercise: 18
Use the values I have obtained for
+
and

to write out equation (149) in full,


using the denitions of Mand Kin equation (145). Expand out the matrix equation
into two ordinary equations, and solve to nd the ratio A
1
/A
2
.
For the frequency
+
, the ratio A
1
/A
2
is -2.73 and for

, the ratio A
1
/A
2
is 0.73. What does all
this mean? Lets go back to the single degree of freedom system, and consider what happened there.
We made a substitution of the form
x = Aexp(it) (160)
in pretty much the same way as we have here, calculated the allowed value of , and that was it.
We couldnt determine the value of A, the amplitude of vibration, without more information. This
extra information came in the form of the initial conditions. For this two degrees of freedom system,
we have found two possible values of , and for each a value of the ratio A
1
/A
2
. Now, from the
substitution we made
x =

Aexp(it) , (161)
we can see that the rst component A
1
of the vector

A represents the amplitude of vibration of
the left mass since it is measured by coordinate x
1
, and the second component A
2
represents the
amplitude of vibration of the right mass. Once again, we cannot deduce the values of both A
1
and
A
2
without knowing the initial conditions. However, we can deduce the ratio of the two, and this is
interesting. For the frequency
+
, we know A
1
/A
2
= -2.73. In this case, the amplitude of the left
mass is more than twice that of the right. The minus sign indicates that the direction of oscillation
of the left mass is opposite to that of the right, i.e. they are in antiphase. For the second frequency

, A
1
/A
2
=0.73, and so both masses are in phase, with the amplitude of vibration of the right mass
a little larger than that of the left one.
We see that the system can oscillate at one of its natural frequencies only at if the amplitudes of
the two masses have the right ratio A
1
/A
2
. We call such an oscillation a fundamental mode of the
system. The fundamental modes are often characterized by vectors

S where the components S


1
and
S
2
are chosen such that they have the right ratio S
1
/S
2
= A
1
/A
2
and the vector has length 1, i.e.,
S
2
1
+S
2
2
= 1. Such a vector is referred to as the mode shape. For our example system, the mode
shapes are for =
+

S
+
=
_
0.94
0.34
_
(162)
and for =

=
_
0.59
0.81
_
(163)
45
Each of
x(t) = Re
_
A
+
_
0.94
0.34
_
exp(i
+
t)
_
(164)
and
x(t) = Re
_
A

_
0.59
0.81
_
exp(i

t)
_
(165)
are valid solutions of the equation of motion. The constants A
+
and A

are in general complex, so


by taking the real part of the complex exponential we get a combination of sines and cosines, or
alternatively a sine with a phase shift.
The general solution can be written as a sum of both solutions,
x(t) = A
+
_
0.94
0.34
_
sin(
+
t +
+
) +A

_
0.59
0.81
_
sin(

t +

) , (166)
where the mode amplitudes A
+
and A

are now real and together with the phases


+
and

must
be determined from initial conditions. So we need four initial conditions. In general, for a system
with n degrees of freedom 2n initial conditions are required.
3.2 Eigenvalue problems
We found in the above section that we could write our equations of motion in matrix form as
M

x +Kx = 0 , (167)
and indeed this is always the case, even for many more degrees of freedom. Using the substitution
x =

Aexp(it) , (168)
we can transform the equation to

2
M

A+K

A = 0 (169)
or rearranging terms
(M
1
K)

A =
2

A . (170)
Now this last form is called the matrix eigenvalue problem, and it arises from many different phys-
ical situations. Because it occurs so often, it has been studied very thoroughly. Moreover, there are
many computer packages that will quickly solve it for us. Lets consider what the solutions are like
rst, then see how a computer can be used to help in the calculations.
If we write
Y =M
1
K (171)
46
and
2
= then the problem we wish to study is
Y

A =

A (172)
The mathematics tells us that if Y is a n n matrix then there are n values of that allow this
equation to have non-trivial solutions, and each of these has associated with it a vector

A. The
values of are called eigenvalues and the associated vectors are eigenvectors.
We have seen that we can take our vibration problem, cast it in matrix form and then manipulate the
matrices so that it is a eigenvalue problem. Using a package such as MATLAB, the solution of this
eigenvalue problem is very easy. In this way, we can tackle systems with quite a large number of
degrees of freedom that would be very hard indeed if we did not have access to a computer.
3.3 Forced vibration
Just as with the single degree of freedomsystems, we would like to be able to calculate what happens
with a forced system. The path from the free body diagram to the equations of motion is no more
complicated than we have met already. Once again, we can write the equations in matrix form so
we end up with something like
M

x +Kx =

F cos(t) , (173)
where the new part is the RHS consisting of a forcing term proportional to cos(t). Many times
only one component of

F will be non-zero, but that need not concern us now. Using what we know
about matrices, solving this really quite difcult problem is surprisingly easy. Just as with the single
degree of freedom systems, we introduce a guessed solution of the form
x =

Acos(t) (174)
where you should note that we have been able to use cos(t) directly rather than the complex
exponential because there is no damping and so we know that the response is either in phase or in
antiphase with the forcing. Substituting this into the equation of motion gives

2
M

Acos(t) +K

Acos(t) =

F cos(t) . (175)
and we can cancel the cosines from each side so we have an equation of the form
Z

A =

F (176)
where
Z =
2
M+K . (177)
The matrix equation (177) corresponds to system of equations for the components of the vector

X.
For small matrices (systems with only a few degrees of freedom) this can be solved by hand. For
47
larger systems, one uses a formal solution which can be computed numerically. To this end, we
have simply to determine the inverse matrix of Z:

A =Z
1

F . (178)
The inverse of a matrix can be calculated using MATLAB. Hence, we may simply use MATLAB to
calculate the inverse matrix Z
1
and multiply this with the vector

F which contains the amplitudes
of the external forces.
Exercise: 19
Go back to the two spring system we dened above and add a forcing term of the
form F
0
cos(t) to the left mass. Use the above method to deduce an expression
for the frequency response curve.
3.4 Orthogonal modes
I want to conclude our consideration of the theory of multiple degree of freedom systems with some
more matrix algebra. Make sure your understanding of matrix algebra is up to scratch before we
start. The following considerations are rather formal, but they lead to a surprising result: A vibrating
system with n degrees of freedom can always be transformed in such a manner that it behaves like
n independent single degree of freedom systems.
The matrix equation of motion for a n degree of freedom system is
M

x +Kx = 0 . (179)
If we now assume that the system is vibrating at one of its fundamental modes, then
x =

S
i
exp(it) , (180)
and we have
K

S
i
=
2
i
M

S
i
, (181)
where we know that there should be n possible values of and Ive acknowledged this in the
equation by labelling these different possible frequencies and the corresponding mode shapes with
an index i which can run from 1 up to n.
Now Im going to prove an important result concerning the vectors

S
i
. Firstly I will multiply the
above equation by the transpose vector

S
T
j
to give me

A
T
j
K

S
i
=
2
i

S
T
j
M

S
i
. (182)
It will become clear later just why I want to do this. For the moment, just make sure you understand
what this means - write out example matrices for yourself if you like. I will also make use of exactly
the same equation written out with the i and j subscripts interchanged,

S
T
i
K

S
j
=
2
j

S
T
i
M

S
j
. (183)
48
Now, the theory of matrix algebra shows that
(AB)
T
=B
T
A
T
. (184)
Using this result, we can also prove quite quickly that
(ABC)
T
=C
T
B
T
A
T
. (185)
If I take my equation (183) above and transpose it, then I get
(

S
T
i
K

S
j
)
T
=
2
j
(

S
T
i
M

S
j
)
T
, (186)
and using the above theorem of matrix algebra gives

S
T
j
K

S
i
=
2
j

S
T
j
M

S
i
, (187)
remembering that the transpose operation twice just goes back to the original matrix and that since
M and K are symmetric the transpose operation has no effect on them. Subtracting equations (182)
and (187) gives me the interesting result
(
2
j

2
i
)

S
T
j
M

S
i
= 0 . (188)
It follows that

S
T
j
M

S
i
= 0 and

S
T
j
K

S
i
= 0 ifi = j . (189)
These equations describe a property of the vectors

S
j
called orthogonality with respect to the matri-
ces M and K. This property turns out to be very useful.
For the case i = j, we dene by

S
T
i
M

S
i
= M
ii
(190)
and

S
T
i
K

S
i
= K
ii
(191)
the generalised mass and stiffness. We will now make use of these ideas in a very powerful method.
Before we can do so, we must make one last denition; an object called the modal matrix P. We
dene it to be the sequence of all the column vectors

S
i
, i.e.
P = [

S
1

S
2
. . .

S
n
] , (192)
so that P is a nn matrix formed by the eigenvectors

S
i
. What is the use of this matrix? Consider
the matrix
P
T
MP = [

S
1

S
2
. . .

S
n
]
T
M[

S
1

S
2
. . .

S
n
] . (193)
Because of the orthogonality relations above, most of the terms when we multiply out this expres-
sion are zero. We are left with
P
T
MP =
_

_
M
11
0 . . . 0
0 M
22
. . . 0
. . .
0 0 . . . M
nn
_

_
(194)
49
which is a diagonal matrix. The matrix P
T
KP is also diagonal. The result we have proved is that
the modal matrix diagonalises the mass and stiffness matrices. This is useful! If we consider the
equation of motion
M

x +Kx =

F
0
exp(it) , (195)
then if we dene a new set of variables byx =Py then
MP

y +KPy =

F
0
cos(t) , (196)
and multiplying this equation by P
T
gives
P
T
MP

y +P
T
KPy =P
T

F
0
cos(t) . (197)
Since the matrices are diagonal, this uncouples the problem: This equation is just n single degree of
freedom problems
M
ii
y
i
+K
ii
y
i
= [P
T

F
0
]
i
cos(t) (198)
which we know how to solve. Hence, our coupled n-degree of freedom system is in fact equivalent
to n uncoupled spring-mass systems. Each of these represents one of the fundamental modes of the
system. The effective force amplitude exciting this mode is given by [P
T

F
0
]
i
=

S
i

F
0
, i.e., it is equal
to the projection of the force vector on the respective mode shape. This is important in practice
since it allows us to assess whether a particular forcing is dangerous in exciting a certain vibration
mode of the system.
The above mathematics seems quite complicated, but dont be intimidated. You can try the method
for yourself. In brief, to solve any forced multiple degree of freedom problem you need to:
Draw a free body diagram
Write down the equation of motion for each component, and put the equations in matrix form
Set the forcing to zero temporarily and solve the freely vibrating systemto nd the frequencies
of free vibration
Use these frequencies to solve the matrix equation and determine the mode shapes
Construct the modal matrix P from the mode shapes
Use the modal matrix to dene a new set of coordinates y
i
(this is conceptual - you dont need
to calculate anything here)
Calculate the products P
T
MP and P
T
KP and write down the new matrix equation for y
Expand out the matrix equation to give n single degree of freedom systems
Solve the single degree of freedom systems
Usex =Py to put the solution back into the original coordinates
50
Its quite a lengthy procedure, but the steps are quite simple to carry out if you have a computer at
hand - and it does give a general method to solve any system, however complicated, provided you
can calculate the eigenvalues and vectors. In particular, the method is suited to programs such as
MATLAB where you can dene and handle matrices easily.
3.5 Applications
3.5.1 Shaft whirling
The next system we will look at is the vibrations of a rotating out-of-balance shaft. For simplicity,
lets consider a massless shaft with a disc of mass m in the centre. Figure 27 shows what this looks
like.
A C e B
G
C e

Figure 27: Schematic diagram of a disc rotating on a shaft. The shaft AB meets the disc at C. The
centre of mass of the disc is located at G.
Figure 28 shows the coordinate system we will use. The origin of the coordinates has been chosen
to be at the equilibrium position of the shaft centre. The shaft centre is attached to the disc at C, so
in equilibrium C rests at O. The centre of mass G is rotated about the shaft centre C at the angular
velocity of the rotation of the shaft, which is . Because the shaft is rotating, C can be displaced
from its equilibrium position and will then be rotated about O. We want to know how far the centre
of the shaft C is displaced from its equilibrium position.
51
v
G
et
e

: C u
O x
Figure 28: Motion of the shaft centre C and the centre of mass G. G moves around C due to the
rotation of the shaft. C moves around O because of the inertial forces caused by rotating G.
Our rst step in analyzing this problem is to draw the free body diagram. This is shown in Figure
29.
v





O x
mee
2
mx G

mv
k: C

Figure 29: Free body diagram for the system of Figure 28, showing the inertial forces and the force
due to the elastricty of the shaft
52
The force acting on C is due to the displacement of the centre of the shaft from its equilibrium
position. If the force for a unit displacement is k then the force here is kz directed towards O. The
value of k is a property of the shaft and the way it is mounted and can be calculated using solid
mechanics. The force acting on G is the centrifugal force due to its rotation about C. This is the
forcing in our problem, which causes C to be displaced from O. The other two forces shown acting
on G are the inertial forces due to its motion. Note that we must give two of these since there are
two independent coordinates needed to specify the position of G. G and C are rigidly linked so the
equation of motion can be deduced by summing all the forces in the system. We get two equations,
by resolving in the x and y directions:
m x +kzcos() me
2
cos(t) = 0 , (199)
m y +kzsin() me
2
sin(t) = 0 . (200)
Now x = zcos() and y = zsin(), so we can write the two equations of motion
m x +kx = me
2
cos(t) , (201)
m y +ky = me
2
sin(t) . (202)
This is a most simple two-degree of freedom system since the two equations are not coupled. There-
fore each of them can be solved using the usual method for forced SDF systems. We nd
x(t) = e

2
/
2
0
1
2
/
2
0
cos(t) , (203)
y(t) = e

2
/
2
0
1
2
/
2
0
sin(t) . (204)
These can be added together to calculate z
2
= x
2
+y
2
. We nd
z = e

2
/
2
0
1
2
/
2
0
. (205)
where
0
is the natural frequency of free vibration for the shaft. If is small, i.e. the shaft ro-
tates slowly, then k m
2
and z = e(/
0
)
2
, which shows C is not far displaced from O. At
higher speeds of rotation, however, C does move a long way from O and at resonance ( =
0
) the
expression goes to innity. At high speeds, for which is large,
z e , (206)
so the centre of mass sits at the origin and the shaft centre C rotates about it.
This simple example shows that shaft whirling occurs at the natural frequency of vibration of the
shaft. This is a very useful result because now we can calculate when whirling will occur by ana-
lyzing the free vibrational properties of the shaft.
53
3.5.2 Beating
We will look at the beating behaviour of two pendulums. This will allow us a bit of practice on
inserting the initial conditions into a MDF system as well as providing an interesting study of the
phenomena of beating.
u
1
u
2
a
k
l
m m
Figure 30: Two weakly coupled pendulums.
u
1
u
2
ka(u
2
u
1
)


mg mg
ml
2
u
1
ml
2
u
2

Figure 31: Free body diagram for the system of Figure 25.
Beating occurs when we have a weak link between two systems with natural frequencies that are
close together. Im going to consider two identical pendulums joined by a weak spring. In reality
the coupling between the pendulums might occur because of the motion of the mounting plate, but
we will just take it as given here, and model it using the weak spring. Figure 30 shows the two
54
pendulums and Figure 31 is the free body diagram.
From the free body diagram, we can write down the equation of motion by taking the moments of
the forces about the pivots of the pendulums.
ml
2

1
+mgl sin
1
ka
2
(
2

1
) = 0 , (207)
ml
2

2
+mgl sin
2
+ka
2
(
2

1
) = 0 . (208)
In the approximation of small oscillations, we can put sin and write the equations of motion
in matrix form:
_
ml
2
0
0 ml
2
__

2
_
+
_
mgl +ka
2
ka
2
ka
2
mgl +ka
2
__

1

2
_
= 0 . (209)
We substitute

=

Aexp(it). If we carry out the differentiation on t and insert the result in the
equation of motion we get
_
ml
2

2
+mgl +ka
2
ka
2
ka
2
ml
2

2
+mgl +ka
2
__
A
1
A
2
_
= 0 . (210)
For the equation to have solutions, the determinant of the matrix must be zero which gives us our
usual quadratic equation for
2
. If we do the algebra we get
ml
2

2
+mgl +ka
2
=ka
2
. (211)
It follows that the values of are

+
=
_
g
l
,

=
_
g
l
+
2ka
2
ml
2
. (212)
Putting these values back into equation (210) gives two equations for the Amplitudes A
1
and A
2
which are
ka
2
_
1 1
1 1
__
A
1
A
2
_
= 0 , =
+
,
ka
2
_
1 1
1 1
__
A
1
A
2
_
= 0 , =

. (213)
We therefore nd that A
1
/A
2
= 1 for =
+
and A
1
/A
2
= 1 for =

. The mode shapes are


therefore

S
+
=
1

2
_
1
1
_
,

S

=
1

2
_
1
1
_
. (214)
Now we will construct the general solution.
The two solutions

S
+
exp(i
+
t) ,

=

exp(i

t) (215)
55
are just the complementary functions. The general solution is just the sum of the complementary
functions multiplied with arbitrary (complex) amplitudes. It is

(t) = Re
_
A
+

2
_
1
1
_
exp(i
+
t) +
A

2
_
1
1
_
exp(i

t)
_
. (216)
At this point in the calculation, I will take the real part. If I dene

ImA
+

2
= B ,
ReA
+

2
=C ,
ImA

2
= D ,
ReA

2
= E , (217)
this gives
_

1

2
_
=
_
Bsin(
+
t) +Ccos(
+
t) +Dsin(

t) +E cos(

t)
Bsin(
+
t) +Ccos(
+
t) Dsin(
+
t) E cos(
+
t)
_
. (218)
This then is the general solution, and it contains four unknowns B,C, D, E that are determined by the
initial conditions. For the initial condition, I will specify that both pendulums are at rest, pendulum
one is at an angle to the vertical and pendulum two is vertical. If we put these conditions into the
above equations for
1
and
2
, we nd C = E = 0 and B = D =/2. Please check this for yourself.
The nal solution is therefore

1
= (/2)[cos(
1
t) cos(
2
t)] (219)
and

2
= (/2)[cos(
1
t) +cos(
2
t)] (220)
We can make use of the trig identities to write this in a more convenient form. We have
cosA+cosB = 2cos
A+B
2
cos
AB
2
cosAcosB = 2sin
A+B
2
sin
AB
2
so our solution can be written as

1
= cos((
1
+
2
)t/2)cos((
1

2
)t/2) (221)
and

2
= sin((
+
+

)t/2)sin((
+

)t/2) (222)
Finally we can put in the values we calculated for
+
and

. If the coupling is very weak, k is


small and so

+
+

2

_
g
l
(223)
56
-1.0
-0.5
0.0
0.5
1.0
2/
2/

2
/

-1.0
-0.5
0.0
0.5
1.0

1
/

Figure 32: Plot of the beating behaviour in the coupled pendulum system. The outer line shows the
sinusoidal envelope curve
57
and

2
=
1
2
_
g
l
_
_
1

1+
2ka
2
mgl
_
_

_
g
l
ka
2
2mgl
(224)
If we write
=
_
g
l
ka
2
2mgl
, =
_
g
l
, (225)
then the solution can be written as

1
= cos(t)cos(t) , (226)

2
= sin(t)sin(t) . (227)
A plot of this form is shown in Figure 32. As you can see, the vibration amplitude is periodically
modulated, and the energy of vibration oscillates between the two pendulums with a period of /.
This phenomenon is called beating.
3.5.3 Anti-resonance and vibration absorbers
We have seen in previous sections that the vibration of a forced system can be always suppressed if
one makes the natural frequency low enough, i.e. by designing a system with large mass and small
stiffness. However, this may often not be a feasible option due to other design constraints.
In this case, another strategy may be chosen to suppress unwanted oscillations. Consider the system
in Figure 33 (left) with a sinusoidal force F(t) = F
0
cos(t). The equation of motion of this system
is
m
1
x
1
+k
1
x
1
= F
0
cos(t) (228)
and the steady-state response of the system is
x(t) =
x
0
1
2
/
2
0
cos(t) =
F
0
k
1
m
1

2
cos(t) . (229)
If the forcing frequency is not much larger than the systems natural frequency
0
=
_
k
1
/m
1
the
response amplitude may be signicant, and at resonance it becomes innite.
How can we suppress the vibration of the mass m
1
without changing the mass or the stiffness of the
system? We will see that we can deal with this problem by adding another mass-spring subsystem
(m
2
and k
2
) as shown in Figure 33 (right). This gives a two-degree of freedom system
m
1
x
1
+k
1
x
1
+k
2
(x
1
x
2
) = F
0
cos(t) (230)
m
2
x
2
+k
2
(x
2
x
1
) = 0 (231)
58
F(t)
x
1
m
1
k
1
x
2
m
2
F(t)
k
2
x
1
m
1

k
1
Figure 33: Left: a forced oscillator. Right: the same oscillator with an additional spring-mass
system acting as anti-vibration device.
or, in matrix form
_
m
1
0
0 m
2
__
x
1
x
2
_
+
_
k
1
+k
2
k
2
k
2
k
2
__
x
1
x
2
_
=
_
F
0
0
_
cos(t) . (232)
We solve this by assuming
_
x
1
x
2
_
=
_
A
1
A
2
_
cos(t) . (233)
where A
1
and A
2
are the vibration amplitudes of the masses m
1
and m
2
. This leads to the system of
equations
(m
1

2
+k
1
+k
2
)A
1
k
2
A
2
= F
0
, (234)
k
2
A
1
+(m
2

2
+k
2
)A
2
= 0 . (235)
By solving this system we obtain the vibration amplitudes A
1
and A
2
. The result is
A
1
=
(k
2
m
2

2
)F
0
K()
, (236)
A
2
=
k
2
F
0
K()
, (237)
59
where
K() = (k
1
+k
2
m
1

2
)(k
2
m
2

2
) k
2
2
. (238)
We see that the amplitude of vibration of the mass m
1
becomes zero if k
2
and m
2
full the relation
k
2
/m
2
=
2
, i.e., if the resonant frequency of the added subsystem matches the excitation frequency.
In particular, we can suppress the resonant response of the original system at =
0
=
_
k
1
/m
1
completely by chosing k
2
/m
2
= k
1
/m
1
.
60
4 Self-excited vibration
We have seen how vibrations can be excited by a force acting from outside on a system. In the
present section, on the other hand, we will be concerned with vibrations that emerge without such
external inuence through a mechanism called positive feedback.
4.1 Positive Feedback
In our vibrating systems we have always been dealing with situations where the damping force
opposes the motion. Positive feedback, on the other hand, means a situation where the motion of
the system leads to a net forward force which accelerates the motion. This leads to a feedback loop:
The accelerated motion increases the forward force, this further accelerates the system motion, and
so on.
To understand how this can happen, we consider the example in Figure 34. This is a fairly common
situation: A slider is moving on a surface, with a friction force that decreases as the velocity of mo-
tion increases. This includes the well-known fact that static friction is larger than kinetic friction,
but also more complicated situations e.g. in the friction of snow and ice: Because of frictional heat-
ing, the motion of a slider creates a water lm which lubricates the sliding interface. This reduces
the friction force the larger the velocity, the larger the heating and the better the lubrication, hence
the lower the friction.
x
c
v m
k F
I
Figure 34: Left: A slider pulled across a surface with a velocity-dependent friction force F
f
and
constant pulling speed y = v
0
.
Let us now analyse a situation where we try to slide a mass m across a surface at a constant velocity
61
y = v
0
. The equation of motion is given by
m x +c( x y) +k(x y) F
f
( x) = 0 . (239)
It is convenient to envisage the motion of the slider in a coordinate system z = x y which moves
at the constant velocity v
0
. Since y = 0, we have
m z +c( z) +kz =F
f
(v
0
+ z) . (240)
If the mass m slides at the constant velocity x = v
0
, then z = 0 and this equation reduces to
kz
0
=F
f
(v
0
) , (241)
i.e., the spring is pulled by such an amount to the right that the spring force offsets the friction force
that has to be overcome for the mass to slide at the imposed velocity v
0
. This is simple, but can such
a situation be sustained?
To investigate this, we assume that the position of the mass m deviates by a small amount z z
0
from the force equilibrium position. We get for z

= z z
0
:
m

+[c +F

f
(v
0
)]

+kz

= 0 . (242)
We see that the damping is replaced by an effective damping [c +F

f
] which depends on the ac-
tual damping constant, but also on the way how the friction force changes with velocity (F

f
is the
derivative of the friction force with respect to velocity, taken at the point v
0
.)
If friction decreases strongly with velocity, [c+F

f
] <0 and the effective damping becomes negative!
Recalling the solution of the free vibration equation, we nd that this implies <0 and the vibration
grows exponentially. That is what we call vibration arising from a positive feedback mechanism.
In machining, several physical mechanisms may lead to positive feedback and self-excited vibration
(chattering vibration). These include
Thermomechanical instability of the cutting process: During cutting, the workpiece is locally
heated and thereby softened. If this effect is pronounced, it may reduce the cutting force and
lead to accelerated cutting which produces even more heat. Ice lubrication as discussed above
may be considered a special case of this thermomechanical feedback mechanism.
Negative velocity sensitivity of friction forces and stick-slip vibration. It is a fairly general
observation even in dry friction that the friction force decreases with increasing velocity.
An example is grinding: If the grinding tool slides faster, this may decrease the experienced
friction force and, if the damping of the tool is not sufciently large, cause further acceleration
of the motion.
62
Regenerative Vibration: Suppose a drill vibrates a bit at its natural frequency. This small
vibration leads to uneven cutting of the workpiece and leaves a pattern on the surface of the
drilled hole. Subsequently the tool cuts into this uneven surface and therefore experiences
a time-dependent force. If the force is such that it increases the vibration, we again have a
positive feedback mechanism that will lead to chattering vibration.
4.2 Stick-slip motion
We analyse in some more detail the dynamical behavior which arises if the friction force decreases
with increasing velocity. In particular, we consider the case of dry friction where the friction force
in Figure 34 can be described by a piecewise constant force law:
|F
f
| =
_

s
F
N
= F
s
, x y = 0 ,

d
F
N
= F
d
, x y = 0 .
(243)
If the slider is at rest, we have static friction with the friction coefcient
s
, otherwise we have
dynamic friction with the friction coefcient
d
<
s
. In either case the friction force is proportional
to the normal force F
N
, and is directed such that it opposes the sliding.
Lets now assume that the slider is initially at rest on the surface, and we start to move the support
to the right at constant velocity y. The force on the mass acting to the left is then c y +kz, and this
is balanced by the static friction force. At a critical elongation z
c
, F
s
= c y +kz
c
, and if the spring is
further elongated, the static friction is overcome and the system starts to move against the surface.
This happens at
z
c
=
F
s
c y
k
. (244)
As soon as the mass starts to slide on the surface, its equation of motion becomes
m z +c z +kz =F
d
. (245)
We know how to solve this equation: We dene z

= z +F
d
/k to get rid of the constant term on the
right-hand side and bring the equation into canonical form. The solution is (see Section 2.3)
z

= Aexp(
0
t)sin(
d
t +) . (246)
The amplitude A and phase follow from the initial conditions at the onset of sliding, which we
identify with the moment t = 0. These are given by
z

(0) = z
c
+
F
d
k
=
F
s
F
d
c y
k
,

z

(0) = y =v
0
. (247)
From these we can work out the amplitude A and phase . For simplicity, we focus on the case
where damping is small, 1. In this case we nd approximately that
d

0
and
z

Aexp(
0
t)sin(
0
t +) ,
z

A
0
exp(
0
t)cos(
0
t +) . (248)
63
Inserting the initial conditions we get
tan =
z(0)
0
v
0

F
s
F
d

0
kv
0
, A =
_
[z

(0)]
2
+[v
0
/]
2
(249)
We nd that the mass undergoes damped oscillation. Now everything depends on whether the
velocity of the mass becomes large enough to allow it to catch up again with the sliding surface.
This occurs if x = z +v
0
= 0, i.e., z =v
0
. As soon as this happens, we switch from sliding to static
friction, i.e., the mass gets stuck and is carried along with the surface. It will then again be released
at the position z
c
, and this stick-slip cycle will repeat itself.
The rst and deepest minimum of z occurs at
0
t + = . Inserting this we nd the condition for
stick-slip to be given by
v
0
< A
0
exp([]) , (250)
which after some transformations can be written as
1 < (1+tan
2
)exp(2[]) . (251)
In the limit of large velocities, large spring stiffnesses, or small friction force difference, 0 and
this condition can never be met. In this case, the mass does not catch up again with the sliding
surface, the oscillation will decay and the system ultimately reaches a regime of steady-state sliding
at the imposed velocity y.
64

Вам также может понравиться