You are on page 1of 46

Lecture Notes

AE2135-II - Vibrations

Version 2.01

Aerospace Structures & Computational Mechanics

c Aerospace Structures and Materials (ASM)


Copyright
All rights reserved.

Table of Contents

III

Table of Contents
1
2

3
4

A
B

C
D
E
F
G

Course overview . . . . . . . . . . . . . . . . . . . . . .
Discretisation . . . . . . . . . . . . . . . . . . . . . . .
2-1
Difference between statics and dynamics . . . . .
2-2
Discretisation . . . . . . . . . . . . . . . . . . .
Free vibrations (undamped) . . . . . . . . . . . . . . . .
Free damped vibrations . . . . . . . . . . . . . . . . . .
4-1
Case 1: Underdamped motion . . . . . . . . . .
4-2
Case 2: Overdamped motion . . . . . . . . . . .
4-3
Case 3: Critically damped motion . . . . . . . .
Forced undamped vibrations (harmonic) . . . . . . . . .
5-1
General solution . . . . . . . . . . . . . . . . . .
5-2
Solution at resonance . . . . . . . . . . . . . . .
Forced damped vibrations (harmonic) . . . . . . . . . .
6-1
Solution method . . . . . . . . . . . . . . . . . .
6-2
Initial conditions . . . . . . . . . . . . . . . . . .
Response to arbitrary excitation . . . . . . . . . . . . . .
7-1
Unit impulse response . . . . . . . . . . . . . . .
7-2
Unit step response . . . . . . . . . . . . . . . . .
7-3
Standard solution . . . . . . . . . . . . . . . . .
7-4
Alternative solution . . . . . . . . . . . . . . . .
Multiple degree of freedom systems . . . . . . . . . . . .
8-1
Solving the MDOF problem . . . . . . . . . . . .
8-2
Modal analysis of MDOF systems . . . . . . . .
8-3
Forced vibrations . . . . . . . . . . . . . . . . .
Simplification standard solution . . . . . . . . . . . . . .
Energy methods . . . . . . . . . . . . . . . . . . . . . .
B-1
Energy analysis . . . . . . . . . . . . . . . . . .
B-2
Euler-Lagrange equation . . . . . . . . . . . . .
Coulomb damping . . . . . . . . . . . . . . . . . . . . .
Logarithmic decrement . . . . . . . . . . . . . . . . . .
Base excitation . . . . . . . . . . . . . . . . . . . . . .
Laplace transforms . . . . . . . . . . . . . . . . . . . . .
General solution strategy . . . . . . . . . . . . . . . . .
G-1
Step 1: DOFs . . . . . . . . . . . . . . . . . . .
G-2
Step 2: Free-body diagram . . . . . . . . . . . .
G-3
Step 3: Equation of motion . . . . . . . . . . . .
G-4
Step 4: Constitutive relations . . . . . . . . . . .
G-5
Step 5: Kinematic relations . . . . . . . . . . . .
G-6
Step 6: Force and moment equilibrium . . . . . .
G-7
Step 7: Combining information into one relation .
G-8
Step 8: Solve for the motion . . . . . . . . . . .

AE2135-II - Vibrations

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

1
2
2
2
7
10
10
12
12
14
14
15
17
17
17
18
18
20
21
23
25
25
28
29
30
32
32
33
34
37
38
39
40
40
40
41
41
41
42
42
42

Lecture Notes

IV

Lecture Notes

Table of Contents

AE2135-II - Vibrations

1 Course overview

Course overview

These notes present a short summary of the content of the course AE2135-II - Vibrations
by Roeland De Breuker. These notes are ultimately based on the book Engineering Vibrations, by Daniel J. Inman. Please feel free to report any mistakes or suggestions to
R.deBreuker@tudelft.nl.
In table 1 you can find the course map for the course, which will bring you to the specified
section when clicked upon.
Table 1: Course map for single degree of freedom systems

Free Vibration

Undamped Motion

Section 3

Forced Vibration
Force + base excitation
Harmonic
Arbitrary
Section 5

Section 7

Section 6

Section 7

Section 4
Damped Motion
Coulomb friction: C
A discussion on multiple degree of freedom systems is given in section 8.

AE2135-II - Vibrations

Lecture Notes

2
2-1

2 Discretisation

Discretisation
Difference between statics and dynamics

There is a difference in the balance of a mass like the one below in a static and a dynamic
case:
Statics: F = 0
Dynamics: F = m
u

u
F2

F3

F1
Figure 1: Balance of a mass

2-2

Discretisation

Continuous structures need to be discretised in order to analyse their dynamic properties


with the tools presented in this course. A possible discretisation of the continuous structure
in figure 2 can be seen in figure 3. The question remains: What are the values of k in the
discretised version? Six different discretisation options are given in the next sections.

w, z
q
x, u
w(x)
L
Figure 2: Continuous beam deformed by a distributed load

F2

F1
k1

w1
m1

k2

w2

k3

m2

L/3

2 L/3
Figure 3: A discretised version of figure 2
Lecture Notes

AE2135-II - Vibrations

2 Discretisation

A: Bending stiffness

F
F
k

E, I, L

Figure 4: Discretising a cantilever beam

In the case of a cantilever beam loaded by a transverse load, the bending stiffness of the beam
is discretised as can be seen in figure 4. For the deflection:
=F

L3
3EI

Hence, in terms of a force-displacement equation, the above can be rewritten as:


F =

F =

3EI
L3

B: Bending stiffness with rotational spring

F
F

kr
E, I, L

Figure 5: Discretising a cantilever beam with a rotational spring

For the case the cantilever beam is attached via a rotational spring (see figure 5), the deflection
is described as follows:
L3
=F
+ L
3EI
Through the moment equilibrium at the root of the beam, the relation between the angle of
rotation and the applied load F can be found:
(

M = kr
M = FL

FL
kr

Hence:
AE2135-II - Vibrations

Lecture Notes

2 Discretisation

F =

F =

3EIkr
L3 kr + 3L2 EI

This equation can be checked by taking the limit of kr . In that case k


the same answer as in the previous case, validating the answer above.

3EI
,
L3

which is

C: Axial stiffness

E, A, L =

Figure 6: Discretising an axially loaded beam

For the axial deflection of the axially loaded beam in figure 6:


=

FL
EA

Hence:
F =

F =

EA
L

D: Axial stiffness of a multi-material rod

E 2 , A 2 , L2

F
E 1 , A 1 , L1

F2 F1

Figure 7: Discretising a parallel multi-material rod

In figure 7, a axially loaded rod composed of two different materials is shown. Its deflection
is calculated using displacement compatibility:
1 = 2 =
Lecture Notes

AE2135-II - Vibrations

2 Discretisation

The right side of the figure shows force equilibrium:


F = F1 + F2
So for the individual internal loads the following can be written:
F1 L
E1 A1
F2 L
2 =
E2 A2
1 =

E1 A1

L
E2 A2
F2 =

L
F1 =

Hence, the force-displacement relation becomes the following:


F =

k


F =

E1 A1 E2 A2
+
L
L

The relation above can again be checked by a limit case, in this case when one of the Youngs
moduli goes to zero. Clearly, the relation can be rewritten to contain the stiffness of the two
individual bars:
F = (k1 + k2 )
E: Axial stiffness of two rods in series

F
E 2 , A 2 , L2
E 1 , A 1 , L1
Figure 8: Discretising two rods in series

There can also be a case where two different types of rod are stacked on top of one another
as displayed in figure 8. In this situation the displacement compatibility as in the previous
example does not count, but is a summation rather:
= 1 + 2
Now the force is constant throughout the rod, and the individual displacements can be described as:
F L1
E1 A1
F L2
2 =
E2 A2

1 =

AE2135-II - Vibrations

Lecture Notes

2 Discretisation

and the total displacement becomes:




=F

L2
L1
+
E1 A1 E2 A2

making the force displacement relation as follows:


F =

F =

L1
E1 A1

1
+

L2
E2 A2

which can be rewritten using the spring constants as in case C (ki =


F =

1
k1

1
+

Ei Ai
Li )

as below:

1
k2

Check the limit case if k1 or k2


F: Rotational stiffness

T,
Figure 9: Discretising for rotational stiffness

For the fixed beam loaded with a torque as displayed in figure 9:


=

TL
GJ

So:

Lecture Notes

T =

kr

T =

GJ
L

AE2135-II - Vibrations

3 Free vibrations (undamped)

Free vibrations (undamped)

The discretised version of the simplified view of a wing as shown in figure 10 is shown on the
right side of the same figure. It is assumed that the mass of the wing is negligible compared
to m.

m
Fuselage

E, I

Fuel pot

L
Figure 10: Discretising a simplified wing

The spring constant has the following value:


3EI
L3

k=
Solution

The free-body diagram (FBD) of the mass is shown in figure 11.

m
mg

FS

+y

Figure 11: Free-body diagram of the mass (the spring is shown for clarity)

The next step is to write down the equation(s) of motion (EOM) and constitutive equation(s)
(CE). The latter are equations that relate forces to displacements.
(

EOM

m
y = Fs + mg

(3.1)

CE

Fs = ky

(3.2)

Combining (3.1) and (3.2) yields an inhomogeneous differential equation:


m
y + ky = mg
To solve this equation, assume a particular solution y p and a homogeneous solution y h :
y = y p + y h = static + dynamic = s + x
AE2135-II - Vibrations

Lecture Notes

3 Free vibrations (undamped)

Regular sine wave

Figure 12: Phase shift of a sine wave

then:
y = x

m
x + k (s + x) = mg

From statics it is known that s =

m
x + kx = mg ks

mg
k

which yields
m
x + kx = 0

(3.3)

x is expected to be harmonic, so assume a solution of the form


x = A sin (n t + )

(3.4)

where A is the amplitude of the oscillation and is the phase shift which describes how
much the sine wave is shifted from x = 0 at t = 0, see figure 12. The standard solution
to the harmonic equation is x = et . How to get from et to A sin (n t + ) is shown in
appendix A. The two unknowns A and can be found from the initial conditions. This is
a second-order homogeneous differential equation, so two initial conditions are needed: the
initial displacement x0 and the initial velocity x 0 . Starting from (3.4), the initial conditions
become:
(

x(0) = x0 = A sin()

(3.5)

x(0)

= x 0 = n A cos()

(3.6)

1: Squaring (3.5) and (3.6), then adding yields


x20 +

x 0
n

2

= A2 sin2 () + cos2 ()
|

{z

A=

=1

x20 +

x 0
n

2

2: Dividing (3.5) by (3.6) gives


x0
1
=
tan()
x 0
n
Lecture Notes

= arctan

x 0 n
x 0

AE2135-II - Vibrations

3 Free vibrations (undamped)

Natural frequencies
Each mass-spring system oscillates at its own natural frequency when excited. This frequency
is n , and is calculated by substituting x = A sin (n t + ) into the homogeneous differential
equation (3.3):
mn2 A sin (n t + ) + kA sin (n t + ) = 0
mn2 = k
s

n =

k
m

Total spring force


Fs = ky
= k (s + x(t))


mg
= k
+ A sin (n t + )
k
The equation above is maximum for

dFs
dt

= 0:

kAn cos (n t + ) = 0

Irrespective of the phase shift , the maximum stays the same, so:

kAn cos (n t) = 0

which holds when cos (n t) = 0, so: n t =

+ k, k Z

t=
n

1
+k
2

so
|Fsmax | = mg + kA

AE2135-II - Vibrations

Lecture Notes

10

4 Free damped vibrations

Free damped vibrations


m
k

Figure 13: A damped mass-spring system

Consider the mass-damper-spring system as shown in figure 13. It has the following equation
of motion:
m
x + cx + kx = 0
To solve for the motion x(t), substitute x = x
et :
t
t
t
m
x
2
e
+ c
x

e
+ k
x

e
=0

m2 + c + k = 0

Solve this with the quadratic formula:


=

c2 4mk
2m

The discriminant c2 4mk makes the difference whether is real or complex. The critical
factor is when it is zero:

c2 = 4mk ccrit = 2 mk
Define the damping ratio :

c
c
=
=
ccrit
2mn
2 mk
Then the equation of motion can be rewritten:
=

x
+ 2n x + n2 x = 0

4-1

Case 1: Underdamped motion

In this case: 0 < < 1:


1,2 =

2n

4 2 n2 4n2
2
q

= n n 2 1
q

= n i n 1 2
|

{z

so:
1 = n id
2 = n + id
Lecture Notes

AE2135-II - Vibrations

4 Free damped vibrations

11

which leads to the following solution (see also figure:




x = en t a1 eid t + a2 eid t

nt
= A |e
{z } sin (d t + )

{z

(4.1)

decay frequency content

Figure 14: Decaying sinusoidal response (the dashed lines indicate en t )

The amplitude A and phaseshift can be found from the initial conditions x(0) = x0 and
x(0)

= x 0 :
x0
x(0) = A sin = x0 A =
(4.2)
sin()
From equation 4.1 the velocity can be deduced:
x = n Aen t sin(d t + ) + Aen t d cos(d t + )
So the initial velocity condition can be written as:
x 0 = n A sin() + Ad cos()

(4.3)

Substituting equation 4.2 in this equation yields:


x 0 = n x0 +
Hence:

= arctan

x0 d
tan()

x 0 d
x 0 + n x0

Also, from equation 4.2 (see figure 15):


sin() =

x0 d
x0
=
A
P

P = Ad

and using the Pythagorean theorem:


q

AE2135-II - Vibrations

A=

x20 d2 + (x 0 + n x0 )2
d

Lecture Notes

12

4 Free damped vibrations

Figure 15: Obtaining P through angle formulae

4-2

Case 2: Overdamped motion

Now: > 1
In this case the roots of the equation of motion are real:
q

1 = n n 2 1
q

2 = n + n 2 1
and the displacement becomes non-oscillatory (see also figure 16):

Figure 16: An overdamped motion with x(0) = 1 and x(0)

=0

4-3

Case 3: Critically damped motion

For this type of motion: = 1


Now the roots are equal:
1 = 2 = n
For repeated roots, the solution takes the form:
x = (a1 + a2 t)en t
where
a1 = x0
a2 = x 0 + n x0
A critically damped motion can be seen in figure
Lecture Notes

AE2135-II - Vibrations

4 Free damped vibrations

13

Figure 17: A critically damped motion with x(0) = 1 and x(0)

= 1

AE2135-II - Vibrations

Lecture Notes

14

5 Forced undamped vibrations (harmonic)

Forced undamped vibrations (harmonic)


F

m
x

Figure 18: A forced mass-spring system

5-1

General solution

EOM:
m
x + kx = F e (t) = F cos(x t)
The displacement x of the forced mass-spring system in figure 18 is composed of a transient
part, xh , and a steady-state part, xp , i.e.:
x = xh + xp
The following is assumed for x:
x = A sin (n t + ) + Ap cos(x t)
|

{z

xh

{z

xp

Solving the steady-state, or particular part of the displacement:


mAp x2 cos(x t) + Ap k cos(x t) = F cos(x t)
so:
Ap =

n2

f
x2

with:

F
f =
m
The amplitude and phase shift are displayed in figure 19.

Lecture Notes

AE2135-II - Vibrations

5 Forced undamped vibrations (harmonic)

15

Figure 19: Amplitude of undamped forced motion and phase difference between loading and
response. After x = n , the load and response are in opposite phase

Suppose:
(

x0 = 0
x 0 = 0

then:
A sin() + Ap = 0
An cos() Ap x 0 = 0

5-2

A = Ap

x = Ap (cos(x t) cos(n t))

+ 2k
2

Solution at resonance
x
+ n2 x = f cos (n t)

The normal assumption to reach a solution would be xp = Ap cos (n t), but in this case this
AE2135-II - Vibrations

Lecture Notes

16

5 Forced undamped vibrations (harmonic)

is not possible as it would lead to Ap . Hence, in this case, we try:


xp = Ap t sin (n t + )
x p = n Ap t cos (n t + ) + Ap sin (n t + )
x
p = n2 Ap t sin (n t + ) + n Ap cos (n t + ) + n Ap cos (n t + )
= 2n Ap cos (n t + ) n2 Ap t sin (n t + )
Substituting this in the equation above yields:
(
(

(
(

p (((((
p (((((
2n Ap cos (n t + ) (
n2(
A(
t sin (n t + ) + (
n2(
A(
t sin (n t + ) = f cos (n t)
2n Ap cos (n t + ) = f cos (n t)

Hence:
= 0 + 2k
f
Ap =
2n
So:

xp =

f
f
t sin (n t + ) =
t sin (n t)
2n
2n

which can be seen in figure 20.

Figure 20: Sinusoidal response, with an amplitude linearly increasing by

Lecture Notes

ft
2n

AE2135-II - Vibrations

6 Forced damped vibrations (harmonic)

17

Forced damped vibrations (harmonic)

Lets look at the harmonic forced vibrations of an underdamped system, this has the following
equation of motion:
x
+ 2n x + n2 x = f cos (x t)

6-1

Solution method

To solve for the displacement,


xp = Re Ap eix t is used, because: eit = cos (t) + i sin (t),

it
so cos (t) = Re e
. Entering this into the EOM gives:




Re

x2 + 2n ix + n2 Ap eix t = Re feix t

f
A = Re
2
2
n x + 2n ix

The denominator can be rewritten as a complex number u = v + iw, with:


v = n2 x2
w = 2n x
Now, denoting u as the complex conjugate of u:
f
fu
A = Re
= Re
u
uu
!

fu
= Re
v 2 + w2

Also:
u = v iw = |u| (cos() i sin()) = |u| (cos() + i sin()) = |u| ei
with = arctan

w
v

. Hence:

!
!

i
2 + w 2 ei
ei
v
f
e
p

A = Re f
= Re
= Re fq
2
2
v 2 + w2
v 2 + w2
2
2
(n x ) + (2n x )

So the particular solution becomes:

fei(x t)

f cos (x t )
= q
xp = Re Ap eix t = Re q
(n2 x2 )2 + (2n x )2
(n2 x2 )2 + (2n x )2


6-2

Initial conditions
x = Ah en t sin (d t + ) + Ap cos (x t )


x = Ah nn t sin (d t + ) + en t d cos (d t + ) Ap x sin (x t )


So when x(0) = 0 and x(0)

= 0, then:
(

Ah sin () + Ap cos () = 0
Ah (n sin () + d cos ()) + Ap x sin ()

from which and Ah can be derived as a function of the known Ap and .


AE2135-II - Vibrations

Lecture Notes

18

7 Response to arbitrary excitation

Response to arbitrary excitation

To be able to solve the response to an arbitrary excitation, first, the solution methods for
a unit impulse and step response need to be known. They are treated first, after which the
standard solution to an arbitrary excitation is treated.

7-1

Unit impulse response

To calculate the response of a system to a unit impulse, first, the unit impulse function
needs to be defined. This function is called the Dirac delta function and has the following
definition (see also figure 21):
(t a) = 0

for

t 6= a

(t a) 6= 0

for

t=a

(t a)dt = 1

has the unit of 1/s.

area = 1

Figure 21: A visualisation of the Dirac delta function

From the definition, it follows that:


Z

f (t)(t a)dt = f (a)

(t a)dt = f (a)

Lets have a look at the impulse response of an underdamped system:

k
c

m
x

Figure 22: A forced damped mass-spring system

The EOM is:


m
x + cx + kx = P (t)
Lecture Notes

AE2135-II - Vibrations

7 Response to arbitrary excitation

19

and the IC are:


(

x(0) = 0
x(0)

=0

To solve the impulse response problem, the impulse is converted to the initial conditions. To
this end, integrate both sides from 0 to :
Z

(m
x + cx + kx) dt =
0

P (t)dt

This can be split into different integrals, looking at each individual part:

lim

mx(0))

= mx(0
+)
m
xdt = lim mx = lim (mx()
0

lim

cxdt
= lim cx = lim (cx() cx(0)) = cx(0+ )
0

lim

kxdt = lim kxt = lim (kx() kx(0)0) = 0


0

lim

P (t)dt = P

x(0+ ) = 0 from the definition of impulse. The impulse can thus be written as an initial
velocity and the equation can be simplified to:
mx(0
+ ) = P

x(0
+) =

P
m

So the EOM becomes:


m
x + cx + kx = 0
with IC:
(

x(0) = 0
x(0)

P
m

Remember, for an underdamped system (see section 4):


x = Aen t sin (d + )
with
q

A=

x20 d2 + (x 0 + n x0 )2

d

x0 d
= arctan
x 0 + n x0


AE2135-II - Vibrations

Lecture Notes

20

7 Response to arbitrary excitation

which for this case become:

= 0 + 2k
A = x 0 = P
d
md

P en t sin ( t)
d
x(t) = md
0

t0
t<0

= P g(t)

where the displacement function to a unit impulse has been labelled g(x) for later convenience.

7-2

Unit step response

The unit step function H, or Heaviside step function, is defined as (see also figure 23):
(

H(t a) =

0
1

for t < a
for t a

H
1

Figure 23: A visualisation of the Heaviside step function

The Heaviside step function is mathematically linked to the Dirac delta function:

(t a) =
H(t a) =

dH(t a)
dt
Zt

( a)d

The response of an underdamped mass-spring system to a unit step at t = a can thus be


found by integrating the response to a unit impulse at t = a. In this case a step starting at
Lecture Notes

AE2135-II - Vibrations

7 Response to arbitrary excitation

21

t = 0 is regarded, so g(x) from section 7-1 can be inserted in the integral:


Zt

G(t) =

g( )d

Zt

Zt

=
0

1 n
e
sin (d ) d
md

1 n
e
sin (d ) H( )d
md

where the lower limit has been set to zero because of the inclusion of the Heaviside function
i t
id t
and we are only interested in the displacement after t = 0. Substitute sin(d t) = e d e
:
2i
Zt

G(t) =
0



1
en eid eid H( )d
2imd
"

1
e(n id )
e(n +id )
=

2imd (n id ) (n + id )

#t
0

The denominator has to be equal to add fractions. Multiplying both fractions with the other
denominator yields a common denominator: (n id ) (n + id ) = 2 n2 + d2 = n2 .
So:
1
2imd n2
1
=
2imd n2
1
=
2imd n2

G(t) =

(n + id )e(n id ) + (n id )e(n +id )




it
0

(n + id ) e(n id )t 1 + (n id ) e(n +id )t 1

n t id t



e
+ n en t eid t


n en t eid t + 
n + id id e
n

+ id id en t eid t

i

h




i
1
n t
id t
id t
id t
id t
2i

e
+
i
e
+
e
n
d
d
2imd n2



n
1
2id
n t
=
1e
2i sin(d t) + 2 cos(d t)
2imd n2
2id
2




1
n
=
1 en t cos(d t) +
sin(d t)
k
d

which is only valid for t 0.

7-3

Standard solution

The response to an arbitrary excitation can be found by dividing the excitation force into
infinitesimally short impulses, see figure 24, and summing all the responses to these impulses
up:
AE2135-II - Vibrations

Lecture Notes

22

7 Response to arbitrary excitation

t
Figure 24: An arbitrary load can be seen as a collection of impulses

P ( ) = f ( )
F ( ) = P ( )(t ) Impulsive force
= f ( ) (t )

linear
system
Figure 25

x( ) = f ( ) g(t )
x(t) = lim

x( ) = lim

f ( ) g(t ) =

Zt

f ( )g(t )d

It does not matter which of the two functions (f or g) is function of which one is function
of (t ), as can be seen below:
t = , so: = t and: d = d
also:

x(t) =

Z0

=0

=t

=t

=0

f (t )g()d =

Zt

f (t )g()d

The response to an arbitrary excitation can be written as:


Zt

x=

f ( )g(t )d

Lecture Notes

AE2135-II - Vibrations

7 Response to arbitrary excitation

23

where g(t) is the impulse response function.


Integrals like these are conveniently solved in the Laplace domain, see also appendix F, where
the integral vanishes into a straightforward multiplication and thus the solution is easily
found. The challenge of this method usually only lies in transforming the solution back from
Laplace domain to the time domain. In the Laplace domain, the equation above becomes:
X = F (s)g(s)
so in case of the following initial conditions:
x(0) = x0
x(0)

= x 0
the equation of motion m
x + cx + kx = f (t) becomes in the Laplace domain:


m s2 X sx0 x 0 + c (sX x0 ) + kX = F (s)


and solving the equation in the Laplace domain can be done by simply isolating X from the
equation:
F (s)
(sm + c)x0 + mx 0
+
+ cs + k
ms2 + cs + k
m [(s + 2n ) x0 + x 0 ]
F (s)
+
=
2
2
m (s + 2n s + n )
m (s2 + 2n s + n2 )
= F (s)g(s) + m [(s + 2n ) x0 + x 0 ] g(s)

X=

ms2

where:
g(s) =

m (s2

1
+ 2n s + n2 )

This expression can be split up in partial fractions and transformed back to the time domain
using a table of standard Laplace transformations (Note: such a table will be provided at the
exam).

7-4

Alternative solution

Instead of using the impulse response function, one can also use the step response function
in the manner as described below. This approach, though slightly more complex, is relevant,
since it is often easier to get step response solutions from numerical codes than impulse
response solutions.
f
H(t )
f =

AE2135-II - Vibrations

x(t) = lim

X f

G(t ) =

Zt

df
G(t )d
d

Lecture Notes

24

7 Response to arbitrary excitation

t
Figure 26: An arbitrary load can also be seen as a collection of steps

Lecture Notes

AE2135-II - Vibrations

8 Multiple degree of freedom systems

25

Multiple degree of freedom systems

An example of a multiple degree of freedom (DOF) system is a wing with two masses, see
figure 27.

Fuselage

m1

k1

m2

k2

Figure 27: A wing with two masses

The wing can basically be discretised in two ways, see figures 28 and 29 below.

k1

m1
k2

x1

m2
x2

Figure 28: Discretising the wing

k1

m1

k2
x1

m2
x2

Figure 29: Discretising the wing in equivalent way

Note that both approaches are equivalent: although for figure 29 the direction of movements
appears to have changed, still wing bending is considered and hence it is just an easier way
to draw the discretised version of the wing.

8-1

Solving the MDOF problem

Upon solving the discretisation method as displayed in figure 29, the loads are as displayed
in figure 30.

AE2135-II - Vibrations

Lecture Notes

26

8 Multiple degree of freedom systems

k1

m1

F1

k2

F2

m2

F2

x1

x2

Figure 30: Finding the loads in the discretised wing

The loads have the following values:


(

F1 = k1 x1
F2 = k2 (x2 x1 )

The equations of motion for the masses are:


(

m1 x
1 = F1 + F2
m2 x
2 = F2

Combining the above:


(

m1 x
1 = k1 x1 + k2 x2 k2 x1
m2 x
2 = k2 x2 + k2 x1

or:
(

m1 x
1 + (k1 + k2 )x1 k2 x2 = 0
m2 x
2 k2 x1 + k2 x2 = 0

This equation can be written in matrix-vector format:


"

m1 0
0 m2

#(

x
1
x
2

"

k + k2 k2
+ 1
k2
k2

#(

x1
x2

( )

0
0

or, in short:
Mx
+ Kx = 0
When handwriting matrix symbols, it is more common to use a double underline, because
writing in bold is not possible.
1. M and K are always symmetric
2. The system is coupled, so you cannot solve x1 without solving x2 and vice versa
The solution can be found by assuming a harmonic displacement again:
x=x
ein t
which stands for:
(

x1 = x
1 ein t
x2 = x
2 ein t

Entering the assumed solution into the matrix equation yields:


("

"

n2 m1
0
k + k2 k2
+ 1
k2
k2
0
n2 m2

|
Lecture Notes

{z

#)

x
=0
}
AE2135-II - Vibrations

8 Multiple degree of freedom systems

27

where A is the system matrix. The trivial solution is x


= 0. A nontrivial solution occurs
when det(A) = 0. As an example, consider:

m1 = 4 103 kg

m = 2 103 kg
2

k1 = 8 kN/m

k2 = 4 kN/m

which yields for the nontrivial solution:




4n2 + 12



2n2 + 4 16 = 0

or:




8 n2 4

n2 1 = 0

so the eigenfrequencies are:


n2 = 4
n2 = 1

n = 2
n = 1

The negative eigenfrequencies can be discarded. Also note that:


s

k1
m1

k2
m2

1 = 1 6=
2 = 2 6=

Entering the value of an eigenfrequency into the equation of motion yields the corresponding
eigenmode. For 1 = 1:
#(
"
)
11
8 4 x
=0
4 2
x
12
The two lines are a scalar multiple of each other, so no unique solution exists. Choose an
equation and choose a value for e.g. x
11 . The first equation yields:

8
x11 4
x12 = 0

Choosing x
11 = 1 gives x
12 = 2 and so the first eigenmode is:
( )

1
2

x
1 =
or any scalar multiple of this. For 2 = 2:
"

#(

4 4
4 4

x
21
x
22

=0

so:
x
21 =
x22
AE2135-II - Vibrations

Lecture Notes

28

8 Multiple degree of freedom systems

-1

(a) The first eigenmode

(b) The second eigenmode


Figure 31

and choosing x
21 = 1 yields x
22 = 1, making the second eigenmode:
(

x
2 =

1
1

or any scalar multiple of this. The two eigenmodes are visualised in figure 31 below. There
are as many eigenmodes and eigenfrequencies as there are degrees of freedom. Considering
the following initial conditions:
n

x1 (0) = x10

x2 (0) = x20 //x 1 (0) = x 10

x 2 (0) = x 20

the complete motion can be written in terms of the two eigenmodes:


x = A1 sin (1 t + 1 ) x
1 + A2 sin (2 t + 2 ) x
2
so:

x1 (t) = A1 sin (1 t + 1 ) x
11 + A2 sin (2 t + 2 ) x
21

x (t) = A cos ( t + ) x
21
1
1 1
1
1 11 + A2 2 cos (2 t + 2 ) x

x2 (t) = A1 sin (1 t + 1 ) x
12 + A2 sin (2 t + 2 ) x
22

x 2 (t) = A1 1 cos (1 t + 1 ) x
12 + A2 2 cos (2 t + 2 ) x
22

and thus, using the initial conditions, there are four equations to solve for the four unknowns:
A1 , A2 , 1 and 2 .

8-2

Modal analysis of MDOF systems

A popular way of solving MDOF systems is by modal analysis. The basic idea is that the
displacements of the MDOF system are replaced by coordinates of the modes (modal coordinates) of the MDOF system. These modes are comparable to the eigenvectors of the MDOF
system as discussed in the previous section. Using the modified eigenvectors of the MDOF
system will result in an uncoupled system of equations which can be solved using the single
degree-of-freedom techniques from the previous chapters in this reader. Obviously from a
physical point-of-view the MDOF system remains coupled, but the coupling is moved for the
system of equations to the modal coordinates.
The procedure is as follows:
1. Mass normalise the displacements x to a new coordinate q.
2. Tranform the mass normalised coordinates q to mass normalised modal coordinates r.
Lecture Notes

AE2135-II - Vibrations

8 Multiple degree of freedom systems

29
1

The mass normalisation is carried out by transforming x = M 2 q. Substituting this into the
equations of motion yields:
1

M M 2 q + KM 2 q = 0
1

Premultiplying the above equation by M 2 yields:


1

M 2 M M 2 q + M 2 KM 2 q = 0
This results in the following equation:
=0
q + Kq
is the mass normalised stiffness matrix, which is still symmetric, but full.
In this equation, K
It is obvious that the eigenvalues of this matrix are the squares of the eigenfrequencies of the
MDOF system.
The next step is to tranform the mass normalised displacements q to modal coordinates r.
This is done by applying the transformation q = P r. Matrix P contains the orthonormal
This transformation results in the following equation of motion:
eigenvectors of K.
r=0
P r + KP
Premultiplying the equation by P T yields:
r=0
P T P r + P T KP
or:
r + r = 0
This system of equaMatrix is a diagonal matrix which contains the eigenvalues of K.
tions is a decoupled system of which the modal amplitudes can be solved individually. The
1
displacement vector can be retrieved by using x = M 2 P r.

8-3

Forced vibrations

Solving MDOF systems with forced vibrations, whether they are harmonic or arbitrary becomes rather straightforward is the decoupled system in modal coordinates is used. The same
methodologies from the previous chapters on single degree-of-freedom systems can be used to
obtain the forced response of r, and the coordinate transformation back to x can be used to
obtain the physical displacements. Obviously the forcing terms have to be modified to:
1

r + r = P T M 2 F
AE2135-II - Vibrations

Lecture Notes

30

A Simplification standard solution

Simplification standard solution

This section shows how to get from x


et to A sin t +
The formal solution to the second order homogeneous differential equation m
x + kx = 0 is
x=x
et


t + k x
t = 0
Substitute this into the differential equation: m2
x
e
e


2 =

k
m

= in

Therefore the solution becomes


x = a1 ein t + a2 ein t
a1 and a2 need to be complex conjugates for x
to be real, so:
a1 = a + ib
a2 = a ib
In this case we first aim for a solution of the form x = A1 sin (n t) + A2 cos (n t). The
following steps are then to be taken after defining the complex conjugates a1 and a2 :
x = a1 ein t + a2 ein t
= (a + ib)ein t + (a ib)ein t
= (a + ib) (cos (n t) + i sin (n t)) + (a ib) (cos (n t) + i sin (n t))
= (a + ib) (cos (n t) + i sin (n t)) + (a ib) (cos (n t) i sin (n t))
= 2a cos (n t) + (2b) sin (n t)

x = A1 cos (n t) + A2 sin (n t)

where A1 = 2a and A2 = 2b. This solution is equivalent to a sine function with a phase
shift. To prove that this is indeed the case, the following identity needs to be used:
ein t = cos (n t) + i sin (n t)
iein t = i cos (n t) sin (n t)
so:


cos (n t) = Im iein t


sin (n t) = Im ein t

and thus the equation for x can be rewritten to:




x = A1 Im iein t + A2 Im ein t


= Im (A1 i + A2 ) ein t
Lecture Notes

AE2135-II - Vibrations

A Simplification standard solution

31

Figure 32: The position of an imaginary number on the imaginary plane

From figure 32 it can be deduced that for an imaginary number A1 i + A2 the following is
valid:


A1
= arctan
A2
so

A1 =
A2 =

A21 + A22 sin()

A21 + A22 cos()

and
(A1 i + A2 ) =

A21 + A22 (cos() + i sin())

= Aei
with A =

A21 + A22 . So now:

AE2135-II - Vibrations

x = Im Aei ein t = Im Aei(n t+)

x = A sin (n t + )

Lecture Notes

32

B Energy methods

Energy methods

Vibration problems can also be solved considering the energy in the system.

B-1

Energy analysis

Equate the potential energy U and kinetic energy T of a system:


U = T

T + U = constant

Tmax = Umax

The kinetic and potential energy are always exchanged to one another, causing the oscillation
around the equilibrium position. Consider the suspended mass-spring system of figure 33:

k
m

Figure 33: A suspended mass-spring system

In this system:
1
Uspring = k ( + x)2
2
Ugravity = mgx
1
T = mx 2
2
Now:

1
1
T + U = constant = k ( + x)2 mgx + mx 2
2
2
Taking the derivative to time yields:

(B.1)

k ( + x) x mg x + mx
x = 0

x + kx) x = 0
(k mg) x + (m
|

{z

=0

{z

=0

where in the last equation, coefficients have been grouped according to their dependency on
x. To find n , a solution of the form x = A sin (t + ) is suggested. In this case:
1
1
Tmax = m (n A)2 = Umax = kA2
2
s2

Lecture Notes

n =

k
m

AE2135-II - Vibrations

B Energy methods

B-2

33

Euler-Lagrange equation

The Euler-Lagrange equation can be used to find equations of motion in a very effective way,
and reads the following:


d L
L
D

+
= Qi
dt x i
xi x i
The Lagrangian L is equal to the kinetic energy minus the potential energy: L = T P ,
D is Rayleighs Dissipation Function, and Qi is a generalised force that is not derivable
from a potential or dissipation function, else it would have been included in one of the other
terms. The potential energy P consists of internal potential energy (U ) like strain energy and
external potential energy (V ), for example due to gravity. For the system in figure 33, there
is no dissipation or any applied force, so the Euler-Lagrange equation becomes:
d
dt

L
x i

L
=0
xi

Using the previously derived kinetic and potential energy for the system under consideration,
the Lagrangian becomes:
1
1
L = T P = k ( + x)2 mgx + mx 2
2
2


Entering this in the Euler-Lagrange equation gives:

AE2135-II - Vibrations

d
(mx)
+ k ( + x) mg = 0
dt
m
x + kx = mg k

Lecture Notes

34

C Coulomb damping

Coulomb damping

Coulomb damping is based on friction instead of the viscous damping as treated e.g. in section
4, see figure 34.

m
N

FS

x
m

Fd

Figure 34: A mass-spring system with friction, the FBD is shown below

The amount of damping is now determined by the friction coefficient . The damping force
depends on the amount of normal force on the surface: Fd = N . The damping coefficient
usually differs for moving or static cases. When there is no movement, friction is governed
by the static friction coefficient s , which is normally higher than the (dynamic) friction
coefficient . For now, however, the friction coefficient is assumed to have only a single
constant value. The equation of motion for the situation in figure 34 consists of two cases:
Suppose the following initial conditions: x(0) = x0 , x(0)

= 0.
1:

x > 0 :

FS

2:

x < 0 :

FS

x
Fd

m
x + kx = Fd

x
Fd

m
x + kx = Fd

Phase 1: x < 0 m
x + kx = Fd
The motion becomes the following:
x = A sin (n t + ) +

Fd
k

and the initial conditions can be written as:


x0 = A sin() +
0 = An cos()

Fd
k

+ 2k
2

So the amplitude becomes:


x0 = A +
Lecture Notes

Fd
k

A = x0

Fd
k
AE2135-II - Vibrations

C Coulomb damping

35

and thus the displacement x(t) can be written as:

Fd
Fd
sin n t +
+
x = x0
k
2
k


Fd
Fd
= x0
cos (n t) +
k
k


This solution is only valid for 0 < t < t1 , of which t1 is the point in time where x changes
sign. To find t1 we have to find the first instance where x = 0:


so t1 =
phase.

n ,

at which x =

2Fd
k

Fd
sin (n t) = 0
k


x = n x0

x0 = x1 , which is a new initial condition for the following

Phase 2: x > 0 m
x + kx = Fd
Now the sign of the damping force changes:
x = A sin (n t + )
Applying the initial condition: x
x1 =

Fd
k

= x1 :

Fd
2Fd
x0 = A sin ( + )
k
k

For the velocity we can deduce the phase shift, because x is again zero at t = t1 :
x = n A cos( + ) = 0

+ 2k
2

and thus the amplitude can be derived:


3Fd
x0 = A
k
making the displacement for t1 < t < t2 :


x = x0

3Fd
Fd
cos(n t)
k
k


To calculate the time t2 at which the velocity is zero again, we equate the derivative to zero
again:


3Fd
2
0 = n
x0 sin(n t) = 0 t2 =
k
n
Finally: x(t2 ) = x0 4Fk d . The displacement can be seen in figure 35. The motion stops as
soon as the amplitude becomes low enough for the spring force not to overcome the frictional
force.

AE2135-II - Vibrations

Lecture Notes

36

C Coulomb damping

Figure 35: The displacement of a mass-spring system with coulomb damping

Lecture Notes

AE2135-II - Vibrations

D Logarithmic decrement

37

Logarithmic decrement

Measure x at time t and time t + T with T = 2d . Then the logarithmic decrement is defined
as:


x(t)
= ln
x(t + T )
So for a generic damped displacement x = Aen t sin (d t + ):

t
n
Ae
sin (d t + )

= ln

n (t+T )

Ae
sin d t + d T +
|{z}
2

Hence:

= ln en T = n T
with T =

2
d ,

so:
= 

n

p
=p
2

1 2
n 1


AE2135-II - Vibrations

4 2 + 2

Lecture Notes

38

E Base excitation

Base excitation
m

m
x1

FS

x2
Figure 36: A base-excited mass-spring system and the corresponding free-body diagram

The displacement of the mass spring system in figure 36 is as follows:


x2 = A2 cos(x t)
where, in this case, x stands for the frequency of the base excitation. The spring force in
the mass-spring system depends on both x1 and x2 , where logically the spring force increases
when the spring increases in length:
Fs = k(x2 x1 )
The equation of motion thus reads:
m
x1 + kx1 = kx2
or:
m
x1 + kx1 = kA2 cos(x t)
The amplitude Ap2 of the particular solution to this differential equation is:
Ap2 =

Lecture Notes

n2 x
2
x
2
=
2
2
n x
1 (x /n )2

AE2135-II - Vibrations

F Laplace transforms

39

Laplace transforms

The definition of the Laplace transform L of a function f (x) is the following:


Z

F (s) =

f (t)est dt

= L (f (t))
The first and second derivative are thus as given below:


L f(t) =

f(t)est dt

= f (t)e

st

i
0

+s

f (t)est dt

= f (0) + sF (s)


L f(t) =

f(t)est dt

= f(t)est

i
0

+s

f(t)est dt

= f(0) sf (0) + s2 F (s)


and also:

L (m
x + cx + kx) = m s2 X sx0 x 0 + c (sX x0 ) + kX

AE2135-II - Vibrations

Lecture Notes

40

G General solution strategy

General solution strategy

This section shows the solution steps to be taken that normally solve a generic problem. As
an example, the structure in figure 37 will be used.

c1
+

k2

k1

c2
b

Figure 37: An example of a vibrations problem

G-1

Step 1: DOFs

Define the independent degrees of freedom and set positive directions. In this case, there is
only one degree of freedom, , and a clockwise rotation is set as positive, as indicated already
in figure 37.

G-2

Step 2: Free-body diagram

Draw a free-body diagram in which all parts that were originally connected are separated and
define the internal and reaction forces, see figure 38.
Hint: if you cut something, like a spring, draw the resulting reaction forces as if they want
to bring the two parts back together, i.e. define tensile as positive. Moreover, draw the
displacements belonging to each internal load and define a positive direction (x1 , x2 and x3
in figure 38).

Lecture Notes

AE2135-II - Vibrations

G General solution strategy

41

Fd1
RA

FS 2

FS 1

+ x1

Fm m
Fd2

+ x2

+ x3

Figure 38: Free-body diagram of figure 37

G-3

Step 3: Equation of motion

Set up the equation(s) of motion, describing the acceleration of each mass as a result of the
forces that act on it. In this case there is one EOM:
m
x3 = Fd2 Fm

G-4

Step 4: Constitutive relations

The next step is to construct the constitutive relations, i.e. the connection between forces
and (derivatives of) displacements, and enter them into the equation of motion. Since in this
case there are four parts that cause a load on the structure due to displacements (two springs
and two dampers), four constitutive relations are to be formulated:
Fd2 = c2 (x(t)
x 3 )
FS1 = k1 x1
Fd1 = c1 x 1
FS2 = k2 x2
Hint: When setting up constitutive relations, be sure to check for a positive displacement
whether the force in the spring/damper becomes positive or negative.

G-5

Step 5: Kinematic relations

Write the displacements defined in the FBD as a function of the degrees of freedom of the
system. These equations are called kinematic relations and there should be one for each
AE2135-II - Vibrations

Lecture Notes

42

G General solution strategy

displacement defined in the FBD. In this case, they are (assuming small angles):
x1 = a
x2 = b
x3 = L

G-6

Step 6: Force and moment equilibrium

To write Fm as a function of the displacements, force and moment equilibrium around A is


to be satisfied. The moment is taken positive in positive -direction, so:
FS1 a Fd1 a FS2 b + Fm L = 0
a
a
b
Fm = FS1 + Fd1 + FS2
L
L
L
a
a
b
= k1 x1 + c1 x 1 + k2 x2
L
L
L

G-7

Step 7: Combining information into one relation

The next step is to combine the four types (i.e. motion, constitutive, kinematic and equilibrium equations) into one relation. Starting from the basic EOM in step 3 and replacing Fd2
and Fm by their equations yields:
a
a
b
m
x3 = c2 (x(t)
x 3 ) k1 x1 c1 x 1 k2 x2
L
L
L
Entering now the kinematic relations into the EOM as well gives:
2
2
2


= c2 x(t)
k1 a c1 a k2 b
L
mL
L
L
L

or, upon rearranging and multiplying by L for convenience:




mL2 + c2 L2 + c1 a2 + k1 a2 + k2 b2 = c2 Lx(t)

| {z }
m

{z
c

{z
k

where m is the equivalent mass, c the equivalent damping and k the equivalent stiffness of
the system.

G-8

Step 8: Solve for the motion

As soon as this point is reached, the problem becomes purely mathematical. The aim is to
use any method suitable to find the solution for the motion, in this case that is (t). Identify
the type of equation and check whether it is consistent with the problem. In this case the
equation describes a forced damped motion because of the and forcing term, which seems
correct because figure 37 shows an applied load and dampers. Now the suitable solution
method can be applied, which in this case is described in section 6.
Lecture Notes

AE2135-II - Vibrations

You might also like