You are on page 1of 60

MECA0029-1 : Theory of vibration

Answers to exam questions

16th of December

Contents

1 Analytical dynamics of discrete systems 2

2subsection.1.1

1.2 Starting from the description of the state of a system of N particles in terms of the equation

given by ( 1.1), describe the different types of kinematic constraints that may exist in

mechanical systems. Define and illustrate the concept of generalized co-ordinates using

examples. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

1.3 Using time-integration of the virtual work principle for N particles in the form given by

the equation 1.2, demonstrate Hamilton’s principle for conservative systems. . . . . . . . . 6

1.4 Explain what means the Hamilton’s principle written in the form ( 1.13). Starting from the

general expressions of T and V in terms of displacement and velocity, deduce the Lagrange

equations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

1.5 Starting from the definition of the kinetic energy for a system of N masses (equation (1.24)),

and using the concept of generalized coordinates, uik (xjk , t) = Uik (q1 , ..., qn , t), explicit the

structure of the kinetic energy. Inserting this into Lagrange’s equations, describe then the

classification of generalized inertia forces. . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

1.6 Describe how to apply the Hamilton’s principle in the case of constrained dynamical systems. 12

2 Undamped vibrations of n-degree-of-freedom systems 14

2.1 Define the different terms that appear into the Lagrange equations in the general case. . . 14

2.2 Introduce the concept of normal modes of vibration. Examine also the case of systems

with a neutrally stable equilibrium configuration. Discuss the orthogonality relationships

of normal modes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2.3 Determine the response of a vibrating system to non-zero initial conditions (including the

case when rigid-body modes -RBM- are present). . . . . . . . . . . . . . . . . . . . . . . . 20


2.4 Define the concept of dynamic influence coefficient. Derive an expression for the spectral

expansion of the FRF matrix. Discuss the case of systems with and without rigid-body

modes. Define the concept of pseudo-resonance. . . . . . . . . . . . . . . . . . . . . . . . . 22

2.5 Calculate the response to external loading through modal expansion. Show how each

normal equation may be integrated numerically. . . . . . . . . . . . . . . . . . . . . . . . . 25

2.6 Describe the effect of truncation of the response in the mode superposition method. Show

how this effect may be partially corrected using the mode acceleration method. . . . . . . 27

2.7 Show how to use the method of additional masses to approximate the response of a system

submitted to ground acceleration excitation. . . . . . . . . . . . . . . . . . . . . . . . . . . 28

2.8 Explain why the concept of effective modal masses is useful when the support of a structure

undergoes a rigid body motion. Define what the different terms represented in equation

(2.76). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

3 Damped vibrations of n-degree-of-freedom systems 29

3.1 Write the normal equations for a damped system. Introduce the assumption of modal

damping and discuss its validity in the case of a lightly damped structure. . . . . . . . . . 29

3.2 Calculate the forced harmonic response of a damped system. Derive the spectral expansion

of the FRF matrix in the general case of low damping. . . . . . . . . . . . . . . . . . . . 32

3.3 Discuss the concept of identification using the force appropriation testing. . . . . . . . . 32

3.4 Explain what are the different terms of the FRF matrix written in the form (see below)

equation ( 3.34). How can this equation be used for modal identification? What is the

information that can be extracted from the identified variables? What is the difference

between a complex mode and a real mode? . . . . . . . . . . . . . . . . . . . . . . . . . . 34

4 Continuous systems 35

4.1 Define the Green’s strain tensor. Apply Hamilton’s principle to a continuous system and

derive the equations of motion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

4.2 Write the Hamilton’s principle in the case of prestressed structures and discuss the case of

structural stability analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

5 The finite element method 40

5.1 Introduce the concept of finite elements using the bar in extension as example. . . . . . . 40
5.2 Apply the finite element method to beam in bending without shear deflection. Extend the

method to the case of a three-dimensional beam element. . . . . . . . . . . . . . . . . . . 43

6 Solution methods for the eigenvalue problem 45

6.1 Describe the concept of reduction of the size of an eigenvalue problem using the static

condesation method. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

6.2 Describe the concept of mechanical impedance and how it can be used for substructuring

a finite element model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

7 Direct time-integration methods 48

7.1 Introduce the general formula for direct integration of first-order systems. Define the

concept of an explicit or implicit integration scheme. . . . . . . . . . . . . . . . . . . . . 48

7.2 Newmark . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

7.2.1 Applying Newmark’s integration formulas (equation(7.5)), describe the Newmark’s

integration algorithm for linear systems. . . . . . . . . . . . . . . . . . . . . . . . . 48

7.2.2 According to Fig. 7.1 and the expressions of amplitude and periodicity error, discuss

the stability and accuracy of the method with respect to the choice of parameters

and time-step. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

7.3 Describe the α-method of Hilber-Hughes-Taylor (HHT) and compare it to Newmark’s

method using Figure 7.2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

7.3.1 Description of the Hilber-Hughes-Taylor (a.k.a alpha-) method . . . . . . . . . . . 50

7.3.2 Comparison . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

7.4 Explain how Newmark’s algorithm can be extended to nonlinear systems of general form

as in Eq. (7.11) using, as a reminder, Newmark’s integration formulas Eq. (7.12). . . . . . 52

8 Introduction to nonlinear systems 54

8.1 Describe the different sources of nonlinearity in mechanical systems. . . . . . . . . . . . . 54

8.2 Explain what the main challenges are in nonlinear dynamics. . . . . . . . . . . . . . . . . 55

8.2.1 No principle of superposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

8.3 Apply the harmonic balance method to the example of the Duffing oscillator to calculate

an approximation of its free response. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57


Theory of Vibration

Introduction
Rules of writing

• Always recall the complete statement of the question as this file should work as a standalone for

the exam.

• After the statement, precise which set of slides are related to the question using the footnote

command.

• Answer in English.

• Use references and labels when speaking about figures.

• Include your graphics in the appropriate subfolders (inside the ’Figs’ folder).

• READBACK (relecture): Please write "EDIT" followed by the modification you do when you

want to modify something in a question (excepted for spelling mistakes). It will help if some people

already started to learn before a correction has been made.

• Send a message to Nicolas Guilluy on facebook when you’re done with your work.

• Deadline: 26th of December

University of Liège -1-


Theory of Vibration 1. Analytical dynamics of discrete systems

1 Analytical dynamics of discrete systems


1.1 Using the notation of the Figure 1.1, define the concept of virtual dis-
placement and write the principle of virtual work for one particle. Gen-
eralize the principle for a system of N particles.1

Figure 1.1: Virtual displacement.

Definition of the virtual displacement δui :

δui = u∗i − ui

where δui is arbitrary for t1 < t < t2 .

Principle of virtual work for one particle:

We suppose that the varied trajectory and the real one both pass through the same points at the ends

of the time interval:

δui (t1 ) = δui (t2 ) = 0

d
(Remark that dt (δui ) = δ u̇i ).

The dynamic equilibrium of the particle of mass m can be written in the form:

müi − Xi = 0 (i = 1, 2, 3)

where the first term represent the inertia force and the second one, external forces. We multiply this

equation by the associated virtual displacement and we sum over the components. The virtual work

expression results:
3
P
(m üi − Xi )·δui = 0.
i=1

which shows that: "The virtual work produced by the forces acting on the particle during a virtual

displacement δui is equal to zero".

1 slide 3-8 and Ref. book p.14-18

University of Liège -2-


Theory of Vibration 1. Analytical dynamics of discrete systems

Generalization for a system of N particles:

Each particle k of the system of N particles with mass mk satisfies the dynamic equilibrium:

mk üik − Xik − Rik = 0 (i = 1, 2, 3 and k = 1, ..., N )

where

• i is the considered space direction;

• k is the particle index;

• Xik are the force components composing the known external resultant force;

• Rik are the components of the unknown resultant reaction forces resulting from the kinematic

constraints imposed on the system.

The virtual displacements for every particle k can be written in the form:
δuik = u∗ik − uik (i = 1, 2, 3)

δuik (t1 ) = δuik (t2 ) = 0 (k = 1, ..., N ),


which verify the kinematic constraints imposed on the system.

The virtual work principle for the system of particles is obtained by projecting the dynamic equilibrium

equations on the virtual displacements and by summing up over the particles:


N X
X 3
(mk üik − Xik − Rik )δuik = 0.
k=1 i=1

However, for a system of N particles (let assume, as an example, a simple system with rigid bonds),

projecting the equations on the kinematically admissible displacement directions consist in summing the

equilibrium equations in the direction of the constraint such that the unknown forces of bonding

disappear.

So, since the reaction forces do vanish when projecting the equations of motion onto kinematically

admissible displacement directions, the virtual work principle for the system of particles takes the form

N P
P 3
(mk üik − Xik )δuik = 0.
k=1 i=1

"The virtual work of the forces effectively applied onto a system of particles is zero with respect to any

kinematically compatible virtual displacement iif the system is in dynamic equilibrium".

1.2 Starting from the description of the state of a system of N particles in


terms of the equation given by ( 1.1), describe the different types of
kinematic constraints that may exist in mechanical systems. Define and
illustrate the concept of generalized co-ordinates using examples.2
2 slide 9-16 and Ref. book p.18-23

University of Liège -3-


Theory of Vibration 1. Analytical dynamics of discrete systems

1.2 Starting from the description of the state of a system of N particles in terms of the equation given

by ( 1.1), describe the different types of kinematic constraints that may exist in mechanical systems.

Define and illustrate the concept of generalized co-ordinates using examples.3

Without kinematic constraints, the state of a system of N particles is completely described by 3N

displacement components uik (also called degrees of freedom). Mathematically, the instantaneous

configuration is given by

ξik (t) = xik + uik (xjk , t) i, j = 1, 2, 3 and k = 1, ..., N (1.1)

where,

• ξik (t) is the instantaneous configuration;

• xik is the reference configuration;

• uik (xjk , t) are the displacement components.

But in most mechanical systems, the particles are submitted to kinematic constraints which restrain

their motion.

Different types of kinematic constraints may exist in mechanical systems. Here they are:

Holonomic constraints:

They are defined by implicit relationships of type: f (ξik , t) = 0. If there is no explicit dependance with

respect to time, the constraints are said to be Scleronomic constraints, otherwise they are rheonomic.

"A holonomic constraint reduces by one the number of degrees of freedom of the system".

Non-holonomic constraints:

They often take the form of differential relations: f (ξ˙ik , ξik , t) = 0 . Such relationships are generally

not integrable. They don’t allow reduction of the number of degrees of freedom of the system.

Concept of generalized displacements:

System of N particles → 3N dofs

R holomic constraints → 3N − R dofs.

The system dynamics is defined in terms of n = 3N − R variables called generalized coordinates or

degrees of freedom (dofs).

uik (xjk , t) = Uik (q1 , ..., qn , t),

3 slide 9-16 and Ref. book p.18-23

University of Liège -4-


Theory of Vibration 1. Analytical dynamics of discrete systems

where the qi are these generalized coordinates. What we show in the red box is that we can express the

displacement components in terms of different variables.

For a system with holonomic constraints only, we may write:

n
X ∂Uik
δuik = δqs
s=1
∂qs

and the virtual work equation becomes

n
 N P
3

P P ∂Uik
(mk üik − Xik ) δqs = 0.
s=1 k=1 i=1 ∂qs

Example: the double pendulum in the 2D plane.

Figure 1.2: Double pendulum in 2D plane.

Here, the system kinematics is described by 4 instantaneous position components: ξ11 , ξ21 , ξ12 , ξ22 .

2 holonomic constraints:

2 2
= l12

ξ11 + ξ21 (Pythagore)
(ξ12 − ξ11 )2 + (ξ22 − ξ21 )2 = l22 (Pythagore)

⇒ 4 − 2 = 2 generalized coordinates: θ1 & θ2 .




 ξ11 = l1 cos θ1
ξ21 = l1 sin θ1

 ξ12 = l1 cos
 θ1 + l2 cos (θ1 + θ2 )
ξ22 = l1 sin θ1 + l2 sin (θ1 + θ2 )

1.3 Using time-integration of the virtual work principle for N particles in


the form given by the equation 1.2, demonstrate Hamilton’s principle
for conservative systems.4
1.3 Using time-integration of the virtual work principle for N particles in the form given by the

4 slide 17-22 and Ref. book p.23-24

University of Liège -5-


Theory of Vibration 1. Analytical dynamics of discrete systems

equation 1.2, demonstrate Hamilton’s principle for conservative systems.5

Hamilton’s principle is no more than a time-integrated form of the virtual work principle for N particles

leads to
Zt2 "X
N X
3
#
(mk üik − Xik )δuik dt = 0 (1.2)
t1 k=1 i=1

where δuik are the arbitrary but compatible virtual displacements which verify the end conditions

δuik (t1 ) = δuik (t2 ) = 0.

The displacements of the system are expressed using the generalized displacements in the form:

uik (xjk , t) = Uik (q1 , ..., qn , t). (1.3)

For a system with holonomic constraints only, we may write:

n
P ∂Uik
δuik = ∂qs δqs
s=1

1) Let assume that the applied forces Xik derive from a potential V:

∂V
Xik = − (1.4)
∂uik

So the virtual work of the forces can be expressed in the form:


N X
3 N X3 n n
X X ∂V X ∂Uik X ∂V
Xik δuik = − δqs = − δqs
∂uik s=1 ∂qs s=1
∂qs
k=1 i=1 i=1
k=1
n
(1.5)
X
= Qs δqs = −δV
s=1

∂V
where Qs = − ∂q s
.

2) The term associated with inertia forces is transformed that way:

d
mk üik δuik = (mk u̇ik δuik ) − mk u̇ik δ u̇ik
dt
(1.6)
d 1
= (mk u̇ik δuik ) − δ( mk u̇ik u̇ik )
dt 2
Owing to the definition of the kinetic energy T of the system,

N 3
1 XX
T = mu̇ik u̇ik (1.7)
2 i=1
k=1

The time-integrated form of the virtual work principle may be rewritten in the form:
" N X
3
#t2 Zt2
X
− mk u̇ik δuik + δ (T − V )dt = 0 (1.8)
k=1 i=1 t1 t1

where the first term is equal to 0 because of the end conditions. Indeed the details of calculation are

presented herebelow:
5 slide 17-22 and Ref. book p.23-24

University of Liège -6-


Theory of Vibration 1. Analytical dynamics of discrete systems

Starting from ( 1.2), with ( 1.4) (1.6),

Zt2 X
N X
3 Zt2 X
N X
3 Zt2 XN X 3
d 1
(mk u̇ik δuik )dt − δ mk u̇ik u̇ik dt − Xik δuik dt = 0,
dt 2
t1 k=1 i=1 k=1 i=1
t1 | {z } t1 |k=1 i=1
{z }
= T = δV

and so the equation ( 1.2) becomes the equation ( 1.8).

In terms of generalized coordinates, considering ( 1.3), we have

n
∂Uik X ∂Uik
u̇ik = + q̇s (1.9)
∂t s=1
∂qs

and therfore,

T = T (q, q̇, t) & V = V (q, t) (1.10)

δq(t1 ) = δq(t2 ) = 0 (1.11)

where q T = [q1 ...qn ].

Hamilton’s principle for a conservative system may be stated in the form:


Rt2
"The real trajectory of the system is such that the integral (T − V )dt remains stationary with respect to
t1

any compatible virtual displacement, arbitrary between both instants t1 t2 but vanishing at the ends of

the interval."
Zt2
δ (T − V )dt = 0
(1.12)
t1

δq(t1 ) = δq(t2 ) = 0

1.4 Explain what means the Hamilton’s principle written in the form
( 1.13). Starting from the general expressions of T and V in terms of
displacement and velocity, deduce the Lagrange equations.6

1.4 Explain what means the Hamilton’s principle written in the form ( 1.13). Starting from the

general expressions of T and V in terms of displacement and velocity, deduce the Lagrange equations.7
Zt2
δ (T − V )dt = 0
(1.13)
t1

δq(t1 ) = δq(t2 ) = 0

Meaning of the Hamilton’s principle:

Let us define a set of generalized coordinates q1 , q2 , . . . , qn , each point in this space corresponds to one

configuration of the system. Then, the time evolution of the system defines a curve in the configuration

space.

6 slide 23-24 and Ref. book p.24-25


7 slide 23-24 and Ref. book p.24-25

University of Liège -7-


Theory of Vibration 1. Analytical dynamics of discrete systems

The integral of the equation (1.14) is called an action integral, as it depends of the path from t1 to t2 ,

independently of the choice of the generalized coordinates qj .


Z t2
I= (T − V) dt (1.14)
t1

In this context, the real trajectory of the system is such that the action integral I, defined in equation

(1.14) remains stationnary (i.e. is minimized), with respect to any compatible virtual displacement δq.

Figure 1.3: Variation of δq.

Those compatible virtual displacements are arbitrary between t1 and t2 , but equals 0 in t1 and t2 .

The word stationnary means that the difference of the action integral is zero to the first order of δq.

Then, the Hamilton principle reduces to the expression of equation (1.15), given that it respects the end

conditions δq(t1 ) = δq(t2 ) = 0.


Z t2
δI = 0 ⇔ δ (T − V) dt = 0 (1.15)
t1

Summary:

• The action I does not depend of the choice of coordinates qj .

• The action I describes the entire motion of the system.

Starting from the general expressions of T and V deduce the Lagrange equations.

The general expression of the kinetic and potential energy respectively, expressed in the generalized

coordinates qs takes the form

T = T (q̇s , qs , t) (1.16)

V = V(qs , t) (1.17)

Then, we derive the variation of kinetic energy from equation 1.16,


n  
X ∂T ∂T
δT = δ q̇s + δqs (1.18)
s=1
∂ q̇s ∂qs

We also derive the variation of potential energy from equation 1.17.


n
X ∂V
δV = δqs (1.19)
s=1
∂qs

University of Liège -8-


Theory of Vibration 1. Analytical dynamics of discrete systems

From Hamilton’s principle expressed as in equation (1.15), we derive successively


Z t2 Z t2
δ (T − V)dt = (δT − δV) dt
t1 t1
n  n
!
Z t2  X
X ∂T ∂T ∂V
= δ q̇s + δqs − δqs dt
t1 s=1
∂ q̇s ∂qs i=1
∂qs
n  n
!
Z t2  X
X ∂T ∂T
= δ q̇s + δqs − (−Qs )δqs dt
t1 s=1
∂ q̇s ∂qs s=1
n 
!
Z t2 
X ∂T ∂T
= + Qs δqs + δ q̇s dt = 0 (1.20)
t1 s=1
∂qs ∂ q̇s

Integrating the second term of (1.20) by parts, we get


   t2  
Z t2  ∂T  ∂T Z t2   
 d ∂T
  
δ q̇ dt = δq − δqs  (1.21)
    
s s
 ∂ q̇s |{z}  ∂ q̇s |{z} dt ∂ q̇
  
t1 t1  s |{z} 
|{z} f 0 |{z} f | {z } f
g g t1 g0

As δqs (t1 ) = δqs (t2 ) = 0, the first term of (1.21) vanishes, and (1.20) reduces to

n 
!
Z t2  
X ∂T d ∂T
+ Qs − δqs dt = 0 (1.22)
t1 s=1
∂qs dt ∂ q̇s

As the displacement δqs is arbitrary, the equation (1.22) reduces for each of the n generalized

coordinates which leads to the motion equation as defined by Lagrange:


 
d ∂T ∂T
− + + Qs = 0 (1.23)
dt ∂ q̇s ∂qs

where the two first terms are the generalized inertia forces associated to qs and the last one is the

generalized (external and internal) forces.

1.5 Starting from the definition of the kinetic energy for a system of N
masses (equation (1.24)), and using the concept of generalized
coordinates, uik (xjk , t) = Uik (q1 , ..., qn , t), explicit the structure of the
kinetic energy. Inserting this into Lagrange’s equations, describe then
the classification of generalized inertia forces.8

1.5 Starting from the definition of the kinetic energy for a system of N masses (equation (1.24)), and

using the concept of generalized coordinates, uik (xjk , t) = Uik (q1 , ..., qn , t), explicit the structure of the

kinetic energy. Inserting this into Lagrange’s equations, describe then the classification of generalized

inertia forces.9
Structure of the kinetic energy

We defined previously the total kinetic energy for a system of N masses as in equation (1.24).

N 3
1 XX
T = mk u̇ik u̇ik (1.24)
2 i=1
k=1
8 slide 25-27 and Ref. book p.27-29
9 slide 25-27 and Ref. book p.27-29

University of Liège -9-


Theory of Vibration 1. Analytical dynamics of discrete systems

Deriving the expression of the generalized coordinates with respect to time, we get

uik = Uik (q1 , . . . , qn , t)


n
∂Uik X ∂Uik
⇔ u̇ik = + q̇s (1.25)
∂t s=1
∂qs

Substituting (1.25) into (1.24), we will get

N 3 n
! n
!
1 XX ∂Uik X ∂Uik ∂Uik X ∂Uik
T (q, q̇, t) = mk + q̇s + q̇s = T0 + T1 + T2 (1.26)
2 i=1
∂t s=1
∂qs ∂t s=1
∂qs
k=1

Where the respective contributions Ti are homogeneous forms of degree i in the generalized velocities q̇s .

1. T0 :
N 3  2
1 XX ∂Uik
T0 = mk = T0 (q, t) (1.27)
2 i=1
∂t
k=1

It represents the transport kinetic energy, of the whole system, as it corresponds to the

situation where all degrees of freedom (qs ) are at rest.

2. T1 :
N X
n X 3   
X ∂Uik ∂Uik
T1 = mk q̇s (1.28)
s=1 k=1 i=1
∂t ∂qs

It represents the mutual kinetic energy.

3. T2 :
n n N 3   
1 XXXX ∂Uik ∂Uik
T2 = mk q̇s q̇r (1.29)
2 s=1 r=1 i=1
∂qs ∂qr
k=1

It represents the relative kinetic energy, as it corresponds to what is left when the explicit

dependence of velocities u̇ik with respect to the time is removed.

Classification of generalized inertia forces

First of all, we must express the terms T1 and T2 with the use of the Euler’s formalism.

Euler’s theorem:

If f (x1 , . . . , xn ) is homogeneous of degree m in the variables (x1 , . . . , xn ), the following equality is

satisfied:
n
X ∂f
xf = mf (1.30)
i=1
∂xi

Then, we can use (1.30) formalism to express T1 and T2

n
X ∂T1
T1 = q̇s (1.31)
s=1
∂ q̇s

n
1 X ∂T2
T2 = q̇s (1.32)
2 s=1 ∂ q̇s

University of Liège -10-


Theory of Vibration 1. Analytical dynamics of discrete systems

If we substitute (1.26) into the inertia’s terms of (1.23), we get


 
 
d ∂T ∂T ∂T0 ∂T1 ∂T2  ∂T0 + ∂T1 + ∂T2 
d  
− + = + + −
dt ∂ q̇s ∂qs ∂qs ∂qs ∂qs dt ∂ q̇s ∂ q̇s
 ∂ q̇s 
|{z}
=0
 
∂T0 ∂T1 ∂T2 d ∂T1 ∂T2
= + + − − (1.33)
∂qs ∂qs ∂qs dt ∂ q̇s ∂ q̇s
"  n
#
∂ 2 T1
 X  
∂(T0 + T1 + T2 ) ∂ ∂T1 d ∂T2
= − + q̇r − (1.34)
∂qs ∂t ∂ q̇s r=1
∂qr ∂ q̇s dt ∂ q̇s

We note 3 types of inertia forces :

1. Transport inertia forces:

We get the transport inertia terms by setting q̇s = 0, then we obtain


 
∂ ∂T1 ∂T0
− + (1.35)
∂t ∂ q̇s ∂qs

2. Relative inertia forces:

Those are the forces obtained by assuming that the constraints do not depend explicitly on time

(i.e. ∂Uik /∂t = 0). Then, the remaining terms are only in T2 .
 
∂T2 d ∂T2
− (1.36)
∂qs dt ∂ q̇s

3. Complimentary inertia forces:

The complementary inertia forces contain the remaining terms:

n
X ∂ 2 T1 ∂T1
Fs = − q̇r + (1.37)
r=1
∂ q̇s ∂qr ∂qs

Considering that f = ∂T1 /∂qs , and that xi = q̇r , we can apply Euler’s formalism:

n
X ∂f ∂T1 X ∂ 2 T1
mf = xi ⇔ = q̇r (1.38)
i
∂xi ∂qs r=1
∂ q̇r ∂qs

Hence, the inertia forces reduce to

n n
∂ 2 T1 ∂ 2 T1
X   X
Fs = q̇r − = q̇r grs (1.39)
r=1
∂qs ∂ q̇r ∂qr ∂ q̇s r=1

We note that the coefficients grs do not depend on the velocities q̇s . They only depend on the

generalized displacements and time. The complementary inertial forces have the nature of

Coriolis or Gyroscopic forces.

1.6 Describe how to apply the Hamilton’s principle in the case of


constrained dynamical systems.10
10 Slide 34-37 and Ref. book p.44-45

University of Liège -11-


Theory of Vibration 1. Analytical dynamics of discrete systems

1.6 Describe how to apply the Hamilton’s principle in the case of constrained dynamical systems.11

Let’s remind that in the case of a system composed of N particles, we have 3N coordinates, such that

ξik (t) = xik + uik (xjk , t) i = 1, 2, 3 (1.40)

And we can express the displacement relatively to the generalized coordinates as follows:

uik (xjk , t) = Uik (q1 , . . . , qn , t) (1.41)

In general, we have, for a constrained system,

fr (ξik , t) = 0 (1.42)

Now, let us suppose that a subset of m kinematic constraints12 isn’t explicitly satisfied by the choice of

the n generalized coordinates qs (m < n).

Then, we can express the constraints for the generalized coordinates:

fr (ξik , t) = fr (xik + Uik (qs , t)) = fr (qs , t) = 0 (1.43)

Inducing that their variation is expressed as in (1.44).


n
X ∂fr
δfr = δqs = 0 ∀r = 1, . . . , m (1.44)
s=1
∂qs

We see that the constraints are verified if the δqs define a motion that is orthogonal to the direction

determined by δfr /δqs in the space of generalized coordinates.

As the derivatives determine the directions of the reaction forces, we can express

∂fr
Rrs = λr (1.45)
∂qs

Where the coefficient λ of (1.45) denotes the intensity of the reaction associated to constraint fr , and is

called a Lagrange multiplier. Hence, substituting (1.45) in (1.44), we get the virtual work of the

reaction forces
m X
X n m
X X ∂fr
Rrs δqs = λr δqs
r=1 s=1 r=1 s=1
∂qs
Xm
= λr δfr (1.46)
r=1

We can add the term (1.46) to the general expression of the virtual work, as obtained by the Hamilton’s

principle, leading to:


n
" m
#
Z t2  
X d ∂T ∂T X ∂fr
− + + Qs + λr δqs dt = 0 (1.47)
t1 s=1
dt ∂ q̇s ∂qs r=1
∂qs
11 Slide 34-37 and Ref. book p.44-45
12 We only consider holonomic constraints here.

University of Liège -12-


Theory of Vibration 1. Analytical dynamics of discrete systems

Where the Lagrange multipliers λr are determined to satisfy the conditions given by Eq. (1.44). In the

end, the Lagrange equations together with the complementary conditions form a system of n + m

equations with n + m unknowns:


   m
 d ∂T − ∂T − Qs − P λr ∂fr = 0

 s = 1, ..., n
dt ∂ q̇s ∂qs r=1 ∂qs (1.48)


 f (q , t) = 0 r = 1, ..., m
r s

University of Liège -13-


Theory of Vibration 2. Undamped vibrations of n-degree-of-freedom systems

2 Undamped vibrations of n-degree-of-freedom systems


2.1 Define the different terms that appear into the Lagrange equations in
the general case.13 Starting from this equation, discuss the concept of
equilibrium configuration. Establish the equations of motion governing
the free vibrations of a system about a stable equilibrium position when
the system does not undergo overall motion. How are the equations of
motion modified when the system is undergoing a steady motion?14

The given system of equations is such as:

∂V ∗
   
d ∂T2 ∂T2 ∂D ∂ ∂T1
dt ∂ q̇s
− ∂qs
= Qs (t) − ∂qs
− ∂ q̇s
+ Fs − ∂t ∂ q̇s
(s = 1, ..., n) (2.1)

where we have:

• Qs (t) , the non-conservative (external) generalised forces,

• V = Vext + V int , the total potential,

• V ∗ = V − T0 , the potential modified by the transport kinetic energy (later, will be called

effective potential energy),

• Fs , the generalised gyroscopic forces,

• D , the dissipation function.

This last is simply the combination of, on the one hand, the Lagrange equation for a non-conservative

system with rheonomic (explicitly timing dependence) constraints written as:


 
d ∂T ∂T ∂V ∂D
− = Qs (t) − − (s = 1, ..., n) (2.2)
dt ∂ q̇s ∂qs ∂qs ∂ q̇s

and on the other hand, the inertia forces which allow to rewrite the left part of the equation (2.2) in the

3 following terms:

1. Transport inertia forces obtained by setting q̇s = 0:


 
∂ ∂T1 ∂T0
− + . (2.3)
∂t ∂ q̇s ∂qs

2. Relative inertia forces obtained by assuming that the constraints do not depend explicitly on
∂Uik

time ∂t =0 :
 
d ∂T2 ∂T2
− + . (2.4)
dt ∂ q̇s ∂qs

13 ch1 slides 25,27,30 and Ref. book p.36


14 ch2 slides 1-15 and Ref. book p.59-65

University of Liège -14-


Theory of Vibration 2. Undamped vibrations of n-degree-of-freedom systems

3. Complementary inertia forces (Coriolis or gyroscopic forces):

n
X
Fs = q̇s grs . (2.5)
r=1

T0 , T1 and T2 are respectively the transport, the mutual and the relative kinetic energy.

Equilibrium configuration

Let us call the equilibrium configuration of a system, a time-independent configuration such as:

qs (t) = qs (0) q̇s (t) = 0 (s = 1, ..., n).

Therefore, such a configuration is reached iff :

V ∗ (q, t) = V ∗ (q)

Qs (t) = 0.

Moreover, let us assume that the non-conservative forces acting on the system are zero when q̇s (t) = 0.

It follows from Lagrange’s equation that an equilibrium position is solution of:

∂V ∗ ∂(V − T0 )
= =0 (s = 1, ..., n), (2.6)
∂qs ∂qs

where the imposed transport kinetic energy T0 is supposed to correspond to a uniform rotation or

translation (independent from the speed!).

Stable equilibrium position when the system doesn’t undergo overall motion

If the system does not undergo overall motion, its kinetic energy reduces to only quadratic term

T = T2 (q̇) in the generalized velocities and the equilibrium position is solution of:

∂V
=0 (s = 1, ..., n). (2.7)
∂qs

Since the potential energy is defined only to a constant, we choose it to be zero at equilibrium i.e.

V (0) = 0 when qs = 0 (s = 1, ..., n). Since the generalised displacements qs represent deviations from

equilibrium, the potential as well as kinetic energy could be expanded in the form of a Taylor series in

the neighbourhood of the equilibrium position:

• Linearization of potential energy

University of Liège -15-


Theory of Vibration 2. Undamped vibrations of n-degree-of-freedom systems

The stiffness matrix K is symmetric and positive definite.

• Linearization of kinetic energy

The mass matrix M is symmetric and positive definite.

By using the two last linearized forms, the Lagrange equations provide the system of equations of

motion governing the free vibrations of a conservative system around a stable equilibrium position as

follow:

Mq̈ + Kq = 0. (2.8)

System undergoing steady motion

Let us now consider the more general case of a system undergoing steady motion (uniform rotational or

translational transport motion). Its equilibrium configuration defined by equation (2.6) corresponds to

∂V
equilibrium between restoring forces of potential origin ( ∂qs
term) and centrifugal forces ( ∂T
∂qs term). It
0

is an equilibrium configuration in the sense that the q̇s , which represent the relative velocities with

respect to the transport motion, vanish although the system itself is not at rest. The following results

will be useful for analysing the stability of systems in steady motion such as rotating systems. Let us

perform successively the linearization of the following terms:

University of Liège -16-


Theory of Vibration 2. Undamped vibrations of n-degree-of-freedom systems

• Linearization of the effective potential energy

n n
∂ 2 T0
 
∗ 1 XX ∗ ∗
V (q) ≈ k qr qs where ksr = ksr − (2.9)
2 s=1 r=1 sr ∂qr ∂qs q=0

• Linearization of the relative kinetic energy

n n
1 XX
T2 (q̇) ≈ msr q̇r q̇s (2.10)
2 s=1 r=1

• Linearization of the coupling kinetic energy

n
X n X
X n
T1 ≈ cs q̇s + fsr q̇s qr (2.11)
s=1 s=1 r=1

∂ 2 T1
   
∂T1
where cs = and fsr =
∂ q̇s q=0 ∂ q̇s ∂qr q=0

/!\ cs and fsr remain constant in a transport motion with uniform speed.

The equilibrium configuration generated by steady motion remains stable as long as the condition

V ∗ ≥ 0 is verified, corresponding to the fact that effective stiffness matrix remains positive definite. In

the neighbourhood of such a configuration, the linearized equations of motion take the form:

n
X

[msr q̈r + (fsr − frs ) q̇r + ksr qr ] = 0 (s = 1, ..., n), (2.12)
r=1

or, in matrix form:

M q̈ + G q̇ + K∗ q = 0. (2.13)

with G the matrix of gyroscopic coupling.

The structure of this equation renders the analysis of systems undergoing overall motion is therefore

more difficult.

2.2 Introduce the concept of normal modes of vibration. Examine also the
case of systems with a neutrally stable equilibrium configuration.
Discuss the orthogonality relationships of normal modes.15

Case of systems with a neutrally stable equilibrium configuration

Another case which deserves attention is that of systems that can exhibit a behaviour corresponding to

a global motion but where either the global velocity is negligible or the additional kinetic energy

generated by the global motion remains constant. The determination of global motion and the analysis

of relative motion are then uncoupled problems, the relative motion occurring about a neutrally stable

(and thus non-unique) equilibrium position as suggested by figure 2.1. Examples of such systems are:

15 Ch2 slide 17-29 and Ref. book p.66-71

University of Liège -17-


Theory of Vibration 2. Undamped vibrations of n-degree-of-freedom systems

• a shaft in torsion and simply supported on its bearings,

• an air-plane in rectilinear flight,

• any structure supported on springs which are very soft compared to its stiffness.

Figure 2.1: Neutrally stable equilibrium position.

In these cases, non-zero displacement modes exist such that such that the associated potential

energy vanishes:
1 T
V (q) = q Kq = 0 for q 6= 0 . (2.14)
2

Solutions (non-trivial ones) comes from the following homogeneous system:

K q=0 dtm|K| = 0 . (2.15)

These solutions, denoted u, are called rigid-body modes. They correspond either to transport motion

or to the existence of a mechanism.

Normal modes of vibration

In order to solve the free vibration linear equations (2.8) let us seek a particular solution such that the

generalized coordinates are governed to a common factor by the same temporal law: q = φ(t) x where x

is a vector of constants, known only to a common scale factor, which governs the shape of the motion.

Motions described by the temporal law are called synchronous since all degrees of freedom follow the

same time function. Substituting a solution of this type yields:

φ̈(t) M x + φ(t) K x = 0. (2.16)

Two cases have to be considered at this point:16

6 0)
1. Systems with a stable equilibrium configuration (dtm|K| =

The matrix K is then non-singular. Equation (2.16) can be put in the form:

φ̈(t)
Kx = − Mx. (2.17)
φ(t)
16 Good to know,M is non-singular by definition.

University of Liège -18-


Theory of Vibration 2. Undamped vibrations of n-degree-of-freedom systems

where Kx and M x do not vanish if one excepts the trivial solution x = 0.

It follows respectively a differential equation and an eigenvalue problem:

φ̈(t)
= λ = ω2 K − ω 2 M x = 0,

− and (2.18)
φ(t)

where λ is a real and positive constant with respect to time. For the differential equation, the

solution is:

φ(t) = α cos ωt + β sin ωt, (2.19)

where ω is called the circular frequency. For the eigenvalue problem (non-trivial solutions and for

r = 1, ..., n):

• ωr2 are called the natural frequencies, root of the algebric equation det(K − ω 2 M ) = 0.

• x(r) represent the synchronous free vibration shapes of the system and are called normal

modes

2. Systems with a neutrally stable equilibrium configuration (dtm|K| = 0)

Remember that this system admit rigid-body displacement modes which are solutions of:

Ku = 0 (2.20)

Hence, as in the previous case, we take the displacement:

q = η(t) u (2.21)

which verify the equation of motion (2.8) if:

η̈(t) M u = 0 (2.22)

and thus, since M is positive definite,

η̈(t) = 0. (2.23)

The time-dependent part of the solution corresponds to a uniformly accelerated motion:

η(t) = γ + δt. (2.24)

The number of rigid-body modes u(i) is equal to the degree of singularity of the stiffness matrix

(i.e. the number of rows which have to be removed to achieve to a non-singular K).

Orthogonality relationships of normal modes

University of Liège -19-


Theory of Vibration 2. Undamped vibrations of n-degree-of-freedom systems

The normal modes verify the orthogonality relationships. Note that these last hold when several

eigensolutions degenerate and give rise to a multiple eigenvalue (which is the case when rigid-body

modes exist). The first relation is:

xT(s) K x(s) = γr δrs , where δrs = Kronecker’s symbol. (2.25)

Its meaning is: the virtual work produced by the inertia forces of mode r in a virtual displacement

described by mode s is zero.

The second relation is:

xT(s) M x(s) = µr δrs , where δrs = Kronecker’s symbol. (2.26)

Its meaning is: the virtual work produced by the elastic forces of mode r in a virtual displacement

described by mode s is zero.

Let us note that very often, the modes are mass normalized (µr = 1) such that:
xT(s) K x(r) = ωr2 δrs
(2.27)
xT(s) M x(r) = δrs

where γr /µr = ωr2 and in matrix form:


Matrix of eigenmodes
XT MX = I z }|
X = [x(1) h ...
{
x(n) ]
where (2.28)
XT KX = Ω2
i
Ω2 = diag ω(1)
2
... 2
ω(n)

2.3 Determine the response of a vibrating system to non-zero initial


conditions (including the case when rigid-body modes -RBM- are
present).17

In the absence of external loading, the response of the system may be determined by superposition of

normal modes. Hence, starting from the initial problem:

M q̈ + K q = 0 with q(0) = q0 , q̇(0) = q̇0 . (2.29)

Once again in this part, we are going to introduce 2 cases:

1. System with a stable equilibrium position

Let us solve the system (2.29) by making the change of generalised coordinates:

17 ch2 slides 32-35 and Ref. book p.77-83

University of Liège -20-


Theory of Vibration 2. Undamped vibrations of n-degree-of-freedom systems

The equations of motion become:

M X η̈ + K X η = 0. (2.30)

Premultiplying by XT and taking into account the orthogonality relationship (equations (2.25)

and (2.26)) leads to the system of n uncoupled equations called the normal equations:

η̈ + Ω2 η = 0 or η̈r + Ω2r ηr = 0 (r = 1, ..., n). (2.31)

The system vibrates as if it were made of n independant oscillators corresponding to its

eigenmodes:

ηr = αr cos ωr t + βr sin ωr t. (2.32)

If we put all this in the q(t) form, the general solution is expressed in the form:

n
X
q(t) = (αs cos ωs t + βs sin ωs t) x(s) (2.33)
s=1

The 2n constants (αs and βs ) are obtained from the initial conditions such as:

It shows that αs and βs ωs are the modal coordinates of q0 and q̇0 respectively.

2. System with neutrally stable equilibrium position (so, including RBM)

When the system possesses m rigid-body modes, the solution found in equation (2.33) for the

response to initial conditions is not applicable, some of the frequencies being zero. However a

similar reasoning based on modal superposition can be followed provided that the contributions of

rigid-body and elastic modes are handled separately. So we express the solution as:

University of Liège -21-


Theory of Vibration 2. Undamped vibrations of n-degree-of-freedom systems

2.4 Define the concept of dynamic influence coefficient. Derive an


expression for the spectral expansion of the FRF matrix. Discuss the
case of systems with and without rigid-body modes. Define the concept
of pseudo-resonance.18

Dynamic influence coefficient

The dynamic influence coefficient hkl (ω 2 ) called also admittance or dynamic flexibility of a system

represents the forced vibration amplitude of the degree of freedom qk for a unitary amplitude harmonic

loading applied on degree of freedom ql at frequency ω. The dynamic influence coefficient is nothing

else than the inverse of the impedance matrix and it is possible to get it by the following steps.

Forced vibration of an n-degrees-of-freedom oscillator is defined as the motion resulting from the

application of a harmonic force with constant amplitude:

M q̈ + K q = s cos ωt, (2.34)

where ω is the excitation frequency and s describes the spatial distribution of the excitation amplitude.

Since the forced response must be synchronous to the excitation, then:

q = x cos ωt. (2.35)

Substituting this solution in the equation of motion, we get:


K − ω2 M x = s


(2.36)
⇔ x = ( K − ω 2 M )−1 s.
| {z }
Impedance matrix

At last, we find the dynamic influence matrix such as it’s equal to the Frequency Response Function

matrix:

H(ω 2 ) = (K − ω 2 M)−1 (2.37)

This last equation highlight the influence of such coefficient in the harmonic domain.
18 Ch2, part 2, slide 4-14 and Ref. book p.83-90

University of Liège -22-


Theory of Vibration 2. Undamped vibrations of n-degree-of-freedom systems

Spectral expansion of the FRF matrix

It is possible to get this spectral expansion solving the equation (K − ω 2 M)x = s by an eigenmode

series expansion:
m
X n−m
X
x= αi ui + βs xs , (2.38)
i=1 s=1

where m are the possible rigid-body modes and αi and βs are (thanks to to the eigenmode

orthogonality (see 2.25 and 2.26)):

uTi s xTs s
αi = − βs = (2.39)
ω 2 µi (ωs2 − ω 2 )µs

So:
m n−m
!
1 X ui uTi X xs xTs
x= − 2 + s (2.40)
ω i=1 µi s=1
(ωs2 − ω 2 )µs

So the frequency response function (FRF) takes the form of:

m n−m
1 X ui uTi X xs xTs
H(ω) = − 2 + (2.41)
ω i=1 µi s=1
(ωs2 − ω 2 )µs

The kl component of the admittance matrix is.

m n−m
2 1 X uki uli X xks xls
Hkl (ω ) = − 2 + 2 − ω 2 )µ
(2.42)
ω i=1 µi s=1
(ωs s

System without rigid-body modes

In case with no rigid body modes (u = 0), the FRF is reduced to:

n
X xks xls
H(ω 2 ) = (2.43)
s=1
(ωs2 − ω 2 )µs

An important property of the principal coefficient Hkk is that are always increasing with excitation

frequency, indeed we can see:


n
dHkk X x2ks
= >0 (2.44)
dω 2 s=1
(ωs2 − ω 2 )2 µs

This property implies that between each two successive eigenfrequencies ωr and ωr+1 there is an

anti-resonance frequency ωrk (ωr < ωrk < ωr+1 ), see figure 2.2. To have the resonances and

anti-resonances we can rewrite the diagonal element Hkk (ω 2 ) in the form of a ratio between two

polynomials in ω 2 of order n for the denominator and of order n − 1 for the numerator, so we have n

resonance frequencies and n − 1 anti-resonances:


  2 
Qn−1 ω
s=1 1− ωsk
Hkk (ω 2 ) = gkk   2  (2.45)
Qn ω
s=1 1− ωs

University of Liège -23-


Theory of Vibration 2. Undamped vibrations of n-degree-of-freedom systems

Figure 2.2: Principal dynamic influence coefficient for a system with no rigid-body mode.

where ωsk are the anti-resonance frequencies and ωs are the resonance frequencies, while gkk is the static

influence coefficient and it is equal to:


n
X x2k(s)
gkk = (2.46)
s=1
ωs2 µs

System with rigid-body modes

In this case, we have m rigid body modes. The principal coefficients are in the form:

m n−m
1 X u2ki X x2ks
Hkk (ω 2 ) = − 2
+ (2.47)
ω i=1 µi s=1
(ωs − ω 2 )µs
2

Its derivative with respect to ω 2 is:

m n−m
dHkk 1 X u2ki X x2ks
2
= + >0 (2.48)
dω ( ω 2 )2 i=1 µi s=1
(ωs − ω 2 )2 µs
2

So when ω 2 → 0 then the slope of Hkk tends to infinity as it is possible to see in figure 2.3. In this case

Figure 2.3: Principal dynamic influence coefficient for a system with rigid-body modes.

we can notice that the system presents n − m resonance frequencies (excluding ωr = 0) and n − m

anti-resonances.

Pseudo-resonance

University of Liège -24-


Theory of Vibration 2. Undamped vibrations of n-degree-of-freedom systems

The response of a system tends to infinity in correspondence of resonance frequencies. However, it is

possible to obtain an harmonic type response of finite amplitude even when the excitation frequency

becomes equal to a resonance frequency ωl and this is possible when the loading produces no work on

eigenmode x(l) , i.e. the external loading verifies the following condition:

xT(l) s = 0 (2.49)

This result is obtained from the following relation:

m n−m
!
1 X ui uTi X xs xTs
x= − 2 + s (2.50)
ω i=1 µi s=1
(ωs2 − ω 2 )µs

2.5 Calculate the response to external loading through modal expansion.


Show how each normal equation may be integrated numerically.19

Let’s consider an undamped n-degree-of-freedom system subjected to an external force p(t). Its

behavior is described by the following equation of motion:

M q̈ + K q = p(t) (2.51)

The initial conditions are:

q(0) = q0 and q̇(0) = q̇0 (2.52)

The eigenmodes xr of the undamped system can be used as a basis to represent the response of the

structure since the normal vibration modes provide a complete set for the expansion of an arbitrary

vector:
n
X
q(t) = ηs xs . (2.53)
s=1

Substituting the equation (2.53) in equation (2.51) and premultiplying by each eigenmode xr we obtain:

xTr M xr η¨r (t) + xTr K xr ηr (t) = xTr p(t) (2.54)


| {z } | {z }
µr γr

By exploiting the orthogonality property of the modes, we get n independent normal equations:

η¨r (t) + ωr2 ηr (t) = φr (t) (r = 1, ..., n) (2.55)

with the modal participation factor of mode r (µr is the modal mass of the rth mode):

xTr p(t)
φr (t) = , (2.56)
µr

19 Ch2, part2, slide 14-20 and Ref. book p.91-94

University of Liège -25-


Theory of Vibration 2. Undamped vibrations of n-degree-of-freedom systems

So our response, thanks to the projection into the modal basis, is reduced to the solution of n uncoupled

single-degree-of-freedom systems.

The projection of the initial conditions in modal representation is obtained by multiplying the two

expressions here below by xTr M:

n
X n
X
q0 = ηs (0)xs and q̇0 = η̇s (0)xs . (2.57)
s=1 s=1

So we obtain:
xTr Mq0 xTr Mq̇ 0
ηs (0) = and η̇s (0) = . (2.58)
µr µr

Each of the normal equations η¨r (t) + ωr2 ηr (t) = φr (t) may be integrated in the form of a convolution

product. Indeed, using the impulsive response approach, the general solution of a normal equation

could be computed in the following way:


Z t
ηr (t) = Ar cos ωr t + Br sin ωr t + φr (τ )h(t − τ )dτ, (2.59)
0

where h(t − τ ) is the impulse response.

The last term of the previous equation is a convolution product known as Duhamel’s integral and

represents the forced response to the external load p(t).

[In details] The convolution product can be integrated in closed form assuming a piecewise linear

variation of the excitation for undamped oscillator submitted to initial conditions. We can perform the

time integration in a recurrent manner. In this iteration manner in the time interval [tn , tn+1 ], we

consider ηr,n and η̇r,n as initial conditions at time tn , while ηr,n+1 and η̇r,n+1 as initial conditions at

time tn+1 = tn + ∆t. So the solution at step tn+1 can be expressed as:
Z tn+1
η̇r,n 1
ηr,n+1 = ηr,n cos ωr ∆t + sin ωr ∆t + φr (τ ) sin ωr (tn+1 − τ )dτ (2.60)
ωr ωr tn

The velocity is instead obtained computing the derivative of the previous equation.
Z tn+1
η̇r,n+1 = −ωr ηr,n sin ωr ∆t + η̇r,n cos ωr ∆t + φr (τ ) cos ωr (tn+1 − τ )dτ (2.61)
tn

Now let us introduce the linearity assumption for the excitation on the time interval and the frequency

parameter α (= ωr ∆t):

sin α sin α
1 − sinα α
       
ηr,n+1 cos α α ηr,n − cos α
α φr,n 1
= +
∆t η̇r,n+1 −α sin α cos α ∆t η̇r,n α sin α + cos α − 1 1 − cos α φr,n+1 ωr2
(2.62)

So these relations allows us to compute the desired response by a time iteration.

University of Liège -26-


Theory of Vibration 2. Undamped vibrations of n-degree-of-freedom systems

2.6 Describe the effect of truncation of the response in the mode


superposition method. Show how this effect may be partially corrected
using the mode acceleration method.20

Generally, when we express the system response to external loading through mode superposition, it

occurs that the number of degrees of freedom is so large that the modal expansion has to be restricted

to a subset of k modes:
k
X
q(t) = ηs (t) xs . (2.63)
s=1

Let us suppose for simplicity that the external load is of type:

p(t) = g φ(t), (2.64)

where g indicates a static load distribution applied with the time-variation law φ(t). For a system

initially at rest, the solution of the normal equation number r is given by:

Z t
ηr (t) = φr (τ )h(t − τ )dτ, (2.65)
0

with:
xTr p(t)
φr (t) = (2.66)
µr

So the normal coordinates can be computed as:

t
xTs g
Z
ηs (t) = sin ωs (t − τ )φ(τ )dτ (2.67)
ωs µs 0

The global convergence of the response can be measured using the M-norm:

k
X
||q||2M = qT Mq = µs ηs2 (t) (2.68)
s=1

From the previous two equations we can get:

k  Z t 2
X (xTs g)2 1
||q||2M = sin ωs (t − τ )φ(τ )dτ (2.69)
µs ω
s=1 | {z } | s 0 {z }
Spatialf actor T emporalf actor

Let us notice that the convergence is governed by two terms, the spatial factor and the temporal

factor.

Spatial factor gives a convergence of quasi-static type. This kind of converges occurs if the applied

load g admits a sufficiently accurate spatial representation on the basis of the k < n retained

eigenmodes. This is equivalent to assuming that the load distribution g must be nearly orthogonal to

the n − k omitted eigenmodes which therefore will not be excited or will be only slightly excited.
20 Ch2, part2, slide 21-27 and Ref. book p.95-98

University of Liège -27-


Theory of Vibration 2. Undamped vibrations of n-degree-of-freedom systems

Temporal factor gives a convergence of spectral type, conditioned by the convergence to zero of the

convolution products in the temporal term when progressing in the eigenspectrum of the system. It

thus depends both on the frequency content of the excitation and on the system eigenspectrum.

It is possible to increase the accuracy of the solution by the modal acceleration method. In this method

indeed only the inertial forces are expressed by modal superposition such as, starting from the equation

of motion rewritten in the form:

K q = p(t) − M q̈, (2.70)

and then, by applying the truncated modal representation to only the inertia forces:
k
X
K q = p(t) − η̈s (t) Mxs , (2.71)
s=1

we obtain the following equation:


k
X
q = K−1 p(t) − η̈s (t) K−1 Mxs , (2.72)
s=1

where ηs (t) verifies the following normal equation:

xTs p(t)
η¨s (t) + ωs2 ηs (t) = (2.73)
µs

Furthermore, taking into account that Kxs = ωs2 Mxs , the solution can thus be rewritten as:
k k
!
X −1
X xs xTs
q(t) = ηs (t)xs + K − p(t) (2.74)
s=1
ω2 µ
s=1 s s
| {z } | {z }
mode displacement solution correction term

Pn x(s) xT
Since K−1 can be rewritten as K−1 = s=1
(s)
ωs2 µs thanks to the spectral expansion, then:
k n
!
X X xs xTs
q(t) = ηs (t)xs + p(t) (2.75)
ω 2 µs
s=1 s=k+1 s
| {z } | {z }
mode displacement solution correction term

In conclusion, we can see that in the modal acceleration method, the mode displacement solution is

complemented with the missing terms from the modal expansion of the static response.

2.7 Show how to use the method of additional masses to approximate the
response of a system submitted to ground acceleration excitation.

TO DO

2.8 Explain why the concept of effective modal masses is useful when the
support of a structure undergoes a rigid body motion. Define what the
different terms represented in equation (2.76).
n1
X (uTs(i) γ(r) )2
mT = mS + (2.76)
r=1
µr

TO DO

University of Liège -28-


Theory of Vibration 3. Damped vibrations of n-degree-of-freedom systems

3 Damped vibrations of n-degree-of-freedom systems


3.1 Write the normal equations for a damped system. Introduce the
assumption of modal damping and discuss its validity in the case of a
lightly damped structure.21

The damped vibration problem is governed by equation:

M q̈ + C q̇ + K q = p(t) (3.1)

Let us consider the associated conservative system associated with the real system. It is governed by

the equations:

M q̈ + K q = 0 (3.2)

2
for which the eigenmodes and eigenfrequencies are noted respectively x(r) and ω0r with r = 1, ..., n.

Since the modes x(r) of the associated conservative system form a complete basis one can, as in the

undamped case, solve the system of equations ( 3.1) through the modal expansion:

n
X
q= ηs (t) x(s) (3.3)
s=1

Normal equation for a damped system:

Making use of the eigenmode orthogonality relationships, we obtain:

n
X βrs 2
η̈r + η̇s + ω0r ηr = φr (t) r = 1, ..., n (3.4)
s=1
µr

with the modal damping coefficients: βrs = xT(r) Cx(s) . As the modal damping coefficients βrs don’t

vanish in general, the modal damping matrix [βrs ] is not diagonal so the normal equations remain

coupled by the damping coefficients.

Modal damping assumption for lightly damped structures:

Let us analyze the free vibrations of a damped system starting from the homogeneous system of

equations:

M q̈ + C q̇ + K q = 0 (3.5)

and let us introduce the assumption that the damping terms are of an order of magnitude lower than

the stiffness and mass terms.

This equation admits a solution of the form:

q = z eλt (3.6)

21 slide 17-25 and reference book p.151-156

University of Liège -29-


Theory of Vibration 3. Damped vibrations of n-degree-of-freedom systems

with the eigenvalues λk and the eigenvectors z(k) solutions of:

(λ2k M + λk C + K) z(k) = 0 (3.7)

Without damping, we would have λk = ±iω0k and z(k) = x(k) , ω0k and x(k) being the eigenmodes and

eigenfrequencies of the associated conservative system:

2
(K − ω0k M) x(k) = 0 (3.8)

If the system is assumed to be lightly damped, it can be supposed that the λk and z(k) differ only

slightly from iω0k and x(k) respectively:

λk = iω0k + ∆λ
(3.9)
z(k) = x(k) + ∆z

Substituting this representation into equation ( 3.7) yields:

2
{(−ω0k + 2iω0k ∆λ + (∆λ)2 )M + (iω0k + ∆λ) C + K}(x(k) + ∆z) = 0 (3.10)

Neglecting next the second-order terms, one obtains:

2
(K − ω0k M) ∆z + (2iω0k M + C)x(k) ∆λ + iω0k C(x(k) + ∆z) ' 0 (3.11)

The light-damping assumption allows to neglect the terms in C∆λ and C∆z. We then obtain the

first-order relationship:
2
(K − ω0k M) ∆z + iω0k (C + 2∆λM)x(k) ' 0 (3.12)

from which, after multiplication by xT(k) , the correction to the eigenvalue may be extracted:

xT(k) (C + 2∆λM)x(k) ' 0

βkk (3.13)
∆λ ' −
2µk
providing the approximate expression:
βkk
λk ' iω0k − (3.14)
2µk

Two important consequences:

• Each eigenvalue correction takes the form of a real negative part and thus transforms each term of

th fundamental solution into a damped ascillatory motion.

• This first-order correction involves only the diagonal damping terms βkk = xT(k) Cx(k) and thus the

influence of the non-diagonal terms is only of second order.

University of Liège -30-


Theory of Vibration 3. Damped vibrations of n-degree-of-freedom systems

Therefore one is authorized to neglect the nondiagonal damping terms without significant alteration of

the eigenspectrum of the lightly damped system.

It is also possible to obtain from equation ( 3.12) the eigenmode correction through a modal expansion

in terms od the undamped system eigenmodes:

n
X
∆z = αs x(s) (3.15)
s=1,s6=k

After substitution and premultiplication by xT(l) for l 6= k and taking into account the orthogonality

relationships,
i ω0k βks
αs = 2 − ω2 ) (3.16)
µs (ω0k 0s

and the resulting eigenmode expression:

n
X βks
z(k) = x(k) + i ω0k 2 2 ) x(s) (3.17)
µs (ω0k − ω0s
s=1,s6=k

2 2
This formula is valid only if the βkl are first-order quantities and if (ω0k − ω0s ) remains finite. It shows

that, as long as the eigenfrequencies of the associated conservative system are well-seperate, it may be

considered that the influence on the eigenmodes of the coupling through damping is of the same order

of magnitude as the damping coefficients βks . It also shows that the eigenmode correction is an

imaginary term, having as its consequence the fact that a free vivbration of the damped sustem is no

longer a synchronous motion of the whole system. Indeed,

q = zeλt = x(k) + Im(∆z) e(iω0k +∆λ)t



(3.18)

and degree of freedom qi is no longer in phase with another degree of freedom qj . For lightly damped

systems with well-seperate eigenvalues, the eigenmode expression ( 3.17) can be approximated by

z(k) ' x(k) and the general solution of the free vibration system takes the form

q ' x(k) e(iω0k +∆λ)t (3.19)

which is consistent with the approximation

xT(k) Cx(k) = βkk δks (3.20)

The discussion above shows that the modal damping assumption is in most cases physically consistent

with the low-damping assumption: when a system is weakly damped and when its eigenfrequencies are

well-seperate, the effect of the cross-damping terms βks , k 6= s on the eigenspectrum can be neglected

according to equations ( 3.14) and ( 3.17).

University of Liège -31-


Theory of Vibration 3. Damped vibrations of n-degree-of-freedom systems

3.2 Calculate the forced harmonic response of a damped system. Derive


the spectral expansion of the FRF matrix in the general case of low
damping. 22

Let us assume harmonic motion with excitation frequency ω

M q̈ + C q̇ + K q = f eiωt (3.21)

and let us assume that the response is limited to the forced term q = z eiωt , so that the excitation and

response amplitudes verify the complex equation:

(K − ω 2 M + i ω C) z = f (3.22)

or

z = H(ω)f (3.23)

where H(ω) is the FRF matrix.

The modal damping assumption allows us to construct the modal expansion of the dynamic influence

coefficient matrix in terms of undamped eigenmodes. Indeed, if the response amplitude is developed in

terms of eigenmodes:
n
X
z= αs x(s) (3.24)
s=1

the orthogonality relationships together with the modal damping assumption provide the coefficients:

xT(r) f
αr = 2 − ω2 + 2 i  ω ω ) (3.25)
µr (ω0s s 0s

Hence the spectral expansion of the dynamic influence coefficient matrix:


n
X 1 x(s) xT(s)
H(ω) = 2 − ω2 + 2 i  ω ω (3.26)
µ ω0s
s=1 s s 0s

and the dynamic influence coefficients can be written as:


n
X 1 xk(s) xTl(s)
Hkl (iω) = 2 − ω2 + 2 i  ω ω ) (3.27)
s=1
(ω0s s 0s µs

Obviously, when the excitation frequency ω tends towards 0, the dynamic influence coefficients converge

to the static influence coefficients gkl , as long as the system does not contain rigid-body modes.

3.3 Discuss the concept of identification using the force appropriation


testing. 23

The most frequent objective of a vibration test is to determine the modal characteristics (eigenmodes

and eigenfrequencies) of the associated conservative system. A natural procedure consists of forcing the
22 slide 33-34 and reference book p.160-162
23 slide 37-39 and reference book p.164-170

University of Liège -32-


Theory of Vibration 3. Damped vibrations of n-degree-of-freedom systems

vibration of the structure in each of its eigenmodes successively, requiring proper tuning of the

frequency and force amplitudes in order to reach the appropriate excitation for each mode.

Phase lag quadrature criterion

When a harmonic vibration test is performed on a damped system, the amplitudes of applied forces and

the response amplitudes at the different points verify the complex relationship:

(K − ω 2 M + i ω C) z = f (3.28)

Extracting a given eigenmode of the associated conservative system through appropriate excitation is

equivalent assuming that: z = x(k) and ω = ω0k .

The equation ( 3.28) becomes:

2
(K − ω0k M + i ω0k C) x(k) = f(k) (3.29)

2
where f(k) is the excitation mode which allows one to achieve the appropriate excitation. Because ω0k

and x(k) are eigensolutions of the associated conservative system, one has:

2
(K − ω0k M) x(k) = 0 (3.30)

and the expression of the excitation which makes it possible to excite eigenmode x(k) at its resonance

frequency results from equation ( 3.29):

f(k) = i ω0k C x(k) (3.31)

This shows that the excitation, when appropriate, is in phase with the dissipation forces and thus has a

phase lag of π/2 with respect to the response.

In order to understand the concept of phase quadrature, let us consider the representation of the

equation of motion in the complex plane. When the excitation is approriate, all degrees of freedom are

in phase and dynamic equilibrium is expressed in the form:

2
(K − ω0k M + i ω0k C) x(k) − f(k) = 0 (3.32)

the complex plane representation of which is given in Figure 3.1.

This shows that when the excitation is appropriate, the elastic forces Kx(k) and the inertia forces

2
ω0k Mx(k) cancel each other as if the system were vibrating in an undamped fashion, while the

excitation force f(k) equilibrates the damping forces which are proportional to the velocity and thus

have a phase advance of π/2.

University of Liège -33-


Theory of Vibration 3. Damped vibrations of n-degree-of-freedom systems

Figure 3.1: Representation of the phase quadrature in the complex plane under appropriate excitation.

The phase quadrature criterion may thus be formulated as follows:

”The structure vibrates according to one of the eigenmodes of the associated conservative system if and

only if all degrees of freedom vibrate synchronously and have a phase lag of π/2 with respect to the

excitation.”

Once the phase quadrature criterion is verified during the experimental test, one may take note of ω0k

and mesure the corresponding eigenshape x(k) .

3.4 Explain what are the different terms of the FRF matrix written in the
form (see below) equation ( 3.34). How can this equation be used for
modal identification? What is the information that can be extracted
from the identified variables? What is the difference between a complex
mode and a real mode? 24

Starting with the definition of the FRF matrix:

H(iω) = (K − ω 2 M + i ω C)−1 (3.33)

It can be found that:


n T T
!
X 1 z(j) z(j) 1 z̄(j) z̄(j)
H(i ω) = + (3.34)
j=1
ρj i ω − λ j ρ̄j i ω − λ̄j

where the bar denotes the complex conjugate and ρj is the squared norm of the eigenvector z(j) .

A dynamic influence coefficient takes the form:

n
!
X Ars(k) Ārs(k)
Hrs (ω) = + (3.35)
i ω − λk i ω − λ̄k
k=1

with
zr(k) zs(k)
Ars(k) = (3.36)
ρk

For the complex mode, the components vary with the lag φ between the components.

24 s.40

University of Liège -34-


Theory of Vibration 4. Continuous systems

4 Continuous systems
4.1 Define the Green’s strain tensor. Apply Hamilton’s principle to a
continuous system and derive the equations of motion.25

Let us consider an elastic body undergoing in time a certain motion measured from the undeformed

configuration. The latter is supposed to be time-invariant, and corresponds to the equilibrium position

of the system in the absence of external forces, figure 4.1.

Figure 4.1: Continuous systems: definitions.

Starting by calculing the deformation at every point of the volume with respect to the reference

configuration, i.e. to express the deformation at a point A of coordinates xi before deformation. Let B

be a second point in the neighbourhood of A and occupying the position xi + dxi in the undeformed

state. After deformation, the points A and B occupy the new positions xi + ui and xi + ui + d(xi + ui ),

respectively (figure 4.2).

Figure 4.2: Strain measures of a segment.

Let us denote ds0 and ds as the lengths of segment AB before and after deformation:
ds20 = dxi dxi = dx21 + dx22 + dx23
(4.1)
ds2 = d(xi + ui ) d(xi + ui )
By taking account of:
∂ui
dui = dxj i = 1, 2, 3 (4.2)
∂xj

then one may write


  
∂ui ∂ui
ds2 = dxi + dxj dxi + dxm
∂xj ∂xm
  (4.3)
∂ui ∂uj ∂um ∂um
= dxi dxi + + + dxi dxj
∂xj ∂xi ∂xi ∂xj
25 slide 2-11 and reference book p.214-224

University of Liège -35-


Theory of Vibration 4. Continuous systems

Let us next define the components of Green’s symmetric strain tensor ij in such a way that the

length increment of segment AB is expressed as:

ds2 − ds20 = 2 ij dxi dxj (4.4)

Hence, according to the previous expressions, the strain components may be written:

 
1 ∂ui ∂uj ∂um ∂um
ij = 2 ∂xj + ∂xi + ∂xi ∂xj (4.5)

In the particular case of linear deformation, it becomes:


 
1 ∂ui ∂uj
ij = + (4.6)
2 ∂xj ∂xi

Displacement variational principle

The displacement variational principle is Hamilton’s principle expressed for a continuous system. Let us

recall that it states that:

Among the feasible trajectories of the system subjected to the restrictive conditions:

δu(t1 ) = δu(t2 ) = 0 (4.7)

at the end of the considered time interval [t1 , t2 ], the real trajectory of the system is the stationary point

of the mechanical action in Lagrange’s and Hamilton’s sense:

Z t2 Z t2
δ Lg [u] dt = δ (T − V) dt = 0 (4.8)
t1 t1

The kinetic energy of the continuous system of figure 4.1 may be evaluated from an integration over

the reference volume (by making use of Einstein’s notation), it takes the form:

Z
1
T(u) = ρ0 u̇i u̇i dV (4.9)
2 V0

The total potentiel energy V results from the summation of the strain energy of the body and the

potential energy of the external forces, supposed conservative:

V = Vint + Vext (4.10)

The strain energy can be expressed by:

Z
Vint = W (ij ) dV (4.11)
V0

University of Liège -36-


Theory of Vibration 4. Continuous systems

and the external potential energy is computed by assuming the existence of body forces: X̄i and

external surface tractions t̄i on the portion Sσ of the surface:


Z Z
Vext = − X̄i (t) ui dV − t̄i (t) ui dS (4.12)
V0 Sσ

The displacement field must verify a priori the kinematic conditions:

ui = ūi (t) on Su , ∀t (4.13)

Derivation of equations of motion

We apply Hamilton’s principle to the continuous system with the assumption that the forces applied to

the system don’t depend on the displacement field ui .

After calculation, one otains:


Z t2 Z t2 Z   
∂uj
δ (T − V) dt = t̄j − ni σij + σim δuj dS
t1 t1 Sσ ∂xm
Z      (4.14)
∂ ∂uj
+ σij + σim − ρ0 üj + X̄j δuj dV dt = 0
V0 ∂xi ∂xm
The displacement variation δuj being arbitrary inside V0 and on Sσ , one obtains the natural conditions

expressing the dynamic equilibrium of the body in the volume and on the surface:
 
∂ ∂uj
σij + σim − ρ0 üj + X̄j = 0 in V0
∂xi ∂xm
  (4.15)
∂uj
tj = ni σij + σim = t̄j on Sσ
∂xm
Linear case

When both rotations and displacements are small, the quadratic term can be neglected:
 
1 ∂ui ∂uj
ij = + (4.16)
2 ∂xj ∂xi

The natural conditions become:


∂σij
− ρ0 üj + X̄j = 0 in V
∂xi (4.17)
tj = ni σij = t̄j on Sσ
The equations above are linearised equations of motion for an elastic body undergoing infinitesimal

displacements and rotations. They express equilibrium in the undeformed state V0 ' V .

4.2 Write the Hamilton’s principle in the case of prestressed structures and
discuss the case of structural stability analysis.26

0
A structure is said to be prestressed when it is submitted to a prescribed fiels of initial stresses σij and

to the associated field of initial strains 0ij , both being independent in time. The analysis of such a


structure consists thus in determining the fields σij and u∗i which are added to this initial state.
26 slide 16-17 and reference book p.227-230

University of Liège -37-


Theory of Vibration 4. Continuous systems

Figure 4.3: Undeformed, prestressed and deformed state.

The notations adopted are summarized by Figure 4.3, and the relationships

ij = 0ij + ∗ij (4.18)

0 ∗
σij = σij + σij (4.19)

express respectively the strain state and the stress state measured in the prestressed configuration V ∗

adopted as reference.

Hamilton’s principle is written:


Z t2
δ (T − V) dt = 0 (4.20)
t1

where the variation operates on the displacement from the initial state, δui = δu∗i , T and V are the

kinetic and potential energies accumulated by the structure when passing from the initial to the

deformed state.

For a prestressed structure, Hamilton’s principle becomes:


Rt
δu∗i t12 (T∗ − V∗int − Vg − V∗ext ) dt = 0
(4.21)
δu∗i (t1 ) = δu∗i (t2 ) =0
with, the additional kinetic energy
Z
∗ 1
T = ρ∗ u̇∗i u̇∗i dV (4.22)
2 V∗

The strain energy of incremental linear strains


Z  
1 ∗(1) ∗(1)
V∗int = Cijkl   dV (4.23)
V∗ 2 ij kl

The geometric strain energy due to prestress


Z
0 ∗(2)
Vg = σij ij dV (4.24)
V ∗

The potential of additional external loads


Z Z
V∗ext =− X̄i∗ u∗i dV − t̄∗i u∗i dS (4.25)
V∗ Sσ

This result shows that the additional displacement field u∗i of a prestressed structure is obtained by a

linear analysis from the prestressed state to the deformed state, with the geometric strain energy due to

initial stress as an additional term.

University of Liège -38-


Theory of Vibration 4. Continuous systems

The theory of the prestressed case forms the basis of structural stability analysis: the latter consists in

computing the prestressing forces applied to the structure which render possible the existence of an

equilibrium under only geometrically linear and nonlinear elastic forces. Hamilton’s principle can then

be reduced to the form:


R t2
δu∗i t1
(V∗int + Vg ) dt = 0 (4.26)

As shown by equation ( 4.21), prestressing modifies the vibration eigenfrequencies, and the limit case of

a vanishing frequency corresponds to the limit stability (T∗ = 0).

University of Liège -39-


Theory of Vibration 5. The finite element method

5 The finite element method


5.1 Introduce the concept of finite elements using the bar in extension as
example.27

Principle: Subdivision of the structure into a finite number of elements of simple geometry with

well-identified structural behaviour (bar, beam, membrane, plate, shell, 3-D solid, ...).

Conditions for the interpolation functions of the displacement field:

• Interpolation in terms of piecewise continuous functions (generally of polynomial type).

• The intensity parameters of these functions (i.e. the generalized coordinates) are local values of

the displacement field.

⇒ If both conditions are strictly satisfied, the approximation obtained is kinematically admissible F.E.

approach in the sens of the Rayleigh-Ritz method.

Bar in extension:

Figure 5.1: Bar in extension.

Let us consider the case of a bar in extension possibly subjected to distributed loads X̄(t). The bar is

first divided into N elements of length l as shown in figure 5.1. We call P 1 and P 2 the loads at the

ends of an element, representing the sum of applied external forces on the nodes and of inner normal

forces originating from interaction with neighbouring elements. Next, the displacement field in the

element is approximated by the linear interpolation:

u(x, t) = u1 (t) φ1 (x) + u2 (t) φ2 (x) (5.1)

where u1 (t), u2 (t) are the connector degrees of freedom, namely the axial displacements at both element

ends, also called nodes.

φ1 (x), φ2 (x) are the shape functions of the element, chosen in such a way that:

27 slide 2-12 and reference book p.363-369

University of Liège -40-


Theory of Vibration 5. The finite element method

u(0, t) = u1 (t) and u(l, t) = u2 (t) (5.2)

If no internal parameter is introduced, they result from a linear interpolation:

x x
φ1 (x) = 1 − and φ2 (x) = (5.3)
l l

We can put equation ( 5.1) in matrix form:

u(x, t) = Φe (x) qe (t) x ∈ element e (5.4)

where

Φe (x) = 1 − xl xl is the shape function matrix of element e


 
 
u1 (t)
qe (t) = is the set of degrees of freedom of element e.
u2 (t)
Element kinetic energy
!
Z l Z l
1 1 1 T
Te = m u̇ dx = q̇Te
2
m ΦTe Φe dx q̇e = q̇ Me q̇e (5.5)
2 0 2 0 2 e

where Me is the elementary (or consistent) mass matrix


 
ml 2 1
Me = (5.6)
6 1 2

Element strain energy


2 !
l l
dφTe dφe
Z  Z
1 du 1 1
Vint,e = EA dx = qTe EA dx qe = qTe Ke qe (5.7)
2 0 dx 2 0 dx dx 2

where Ke is the elementary stiffness matrix


 
EA 1 −1
Ke = (5.8)
l −1 1

Virtual work of external forces

∂Vext,e = −∂qTe Pe (t) (5.9)

with the generalized loads pe (t) conjugated to displacements qe (t):

Z l  
P1 (t)
pe (t) = ΦTe X̄(x, t) dx + (5.10)
0 P2 (t)

For a uniform load X̄0 per unit length:


   
X̄0 l 1 P (t)
pe (t) = + 1 (5.11)
2 1 P2 (t)

We verify that:

University of Liège -41-


Theory of Vibration 5. The finite element method
 
• the rigid body mode uTe = 1 1 is such that Ke ue = 0,

• the mass associated to the rigid body mode, obtained by projecting the mass matrix on it, yields

the total mass of the element: uTe Me ue = ml.

Hamilton’s principle applied to the single element e

By using the equations ( 5.5), ( 5.7) and ( 5.9), we obtain:


Z t2 Z t2   Z t2
1 T 1
δ (T − V) dt = δ q̇ Me q̇e − qTe Ke qe dt + δ qTe pe dt = 0 (5.12)
t1 t1 2 e 2 t1

If we apply the virtual variation to the generalized coordinates q, we arrive at the equations of motion

in the discrete form (equilibrium equation) at element level:

Me q̈e + Ke qe = pe (5.13)

where pe contains the a priori unknown reaction forces between element and its neighbours.

Assembly process

In order to express dynamic equilibrium for the global system of figure 5.1, let us construct the matrix

of structural displacements q collecting (N + 1) nodal displacements.

qT = u0
 
u1 u2 ... uN (5.14)

such that the degrees of freedom of an element are retrieved as:

qe = Le q (5.15)

where the localization operator Le is a Boolean matrix with dimension 2x(N + 1). It contains only 1

and 0 terms and can be seen as a filter picking the nodal displacements qe of an element out of the

global set of degrees of freedom q.

By summing all the elements of the system, the structural variational equation becomes:
Z N 
t2 X  Z N
t2 X
1 T 1
δ q̇e Me q̇e − qTe Ke qe dt + δ qTe pe dt = 0 (5.16)
t1 e=1
2 2 t1 e=1

and it may be expressed in terms of structural displacements using equation ( 5.15):


Z t2 " N
! N
! # Z t2 N
!
1 T X T 1 T X T T
X
δ q̇ Le Me Le q̇ − q Le Ke Le q dt + δq Le pe dt = 0
t1 2 e=1
2 e=1 t1 e=1
(5.17)

We define the structural mass matrix M, the structural stiffness matrix K and the structural load

vector p.
Z t2   Z t2
1 T 1
δ q̇ M q̇ − qT K q dt + δ qT p dt = 0 (5.18)
t1 2 2 t1

University of Liège -42-


Theory of Vibration 5. The finite element method

By taking the variation of this final discretized expression, one obtains the discretized structural

equations in the usual form:

M q̈ + K q = p(t) (5.19)

5.2 Apply the finite element method to beam in bending without shear
deflection. Extend the method to the case of a three-dimensional beam
element.28

We consider the case of a beam in bending (figure 5.2), possibly excited by a distributed load p̄(x, t).

Figure 5.2: Beam in bending.

In the context of the kinematic Bernoulli assumptions for beams with no shear deflection, let us express

the strain energy of the beam element (with no prestress) in the form:
L 2
∂2w
Z 
1
Vint,e = EI dx (5.20)
2 0 ∂x2
x
In terms of nondimensional variable ξ = l over the element domain, the approximation of the

displacement field in the element is given by:


w(ξ, t) = w1 φ1 (ξ) + ψ1 φ2 (ξ) + w2 φ3 (ξ) + ψ2 φ4 (ξ)
(5.21)
= Φe (ξ) qe (t)
∂w
ψ= (5.22)
∂x

where the shape functions φi (ξ), (i = 1 ... 4) are the third-order Hermitian polynomials, matching the

kinematic conditions:

dφi
φ1 (0) = 1 φ˙1 (0) = 0 φ1 (1) = φ̇1 (1) = 0 where φ̇i =








 φ2 (0) = 0 φ˙2 (0) = l φ2 (1) = φ̇1 (1) = 0 (5.23)





φ3 (ξ) = φ(1 − ξ) φ4 (ξ) = −φ2 (1 − ξ) = 0

We get the matrix of shape functions:


1 − 3 ξ2 + 2 ξ3
 
 l ξ (1 − ξ)2 
ΦTe (ξ) = 
 ξ 2 (3 − 2 ξ) 
 (5.24)
l ξ 2 (ξ − 1)
28 slide 20-26 and reference book p.376-379

University of Liège -43-


Theory of Vibration 5. The finite element method

associated with the element degrees of freedom:

qTe = w1
 
ψ1 w2 ψ2 (5.25)

The kinetic energy of the element (neglecting rotational energy):


!
Z l Z l
1 1
Te = ρ A ẇ2 dx = q̇Te l m(ξ) ΦTe Φe dξ q̇e
2 0 2 0 (5.26)
1
= q̇Te Me q̇e
2
where Me is the elementary mass matrix.

The strain energy of the element;


2 T  !
l l
∂2w d2 Φe d 2 Φe
Z  Z  
1 dξ 1 1
Vint,e = EI = qTe E I(ξ) dξ qe
2 0 ∂ξ 2 l3 2 l3 0 dξ 2 dξ 2
(5.27)
1
= qTe Ke qe
2
where Ke is the elementary stiffness matrix.

The virtual work of external loads:

δVext,e = −δqTe pe (t) (5.28)

with the vector of external loads


Z l
pe (t) = l ΦTe (ξ) p̄(ξ, t) dξ (5.29)
0

For an element of uniforme characteristics excited by a constant distributed load p̄0 , we explicitly

obtain:  
12 6l −12 6l
EI  6l 4l2 −6l 2l2 
Ke = 3   (5.30)
l  −12 −6l 12 −6l
6l 2l2 −6l 4l2
 
156 22l 54 −13l
ml  22l 4l2 13l −3l2 
Me =  (5.31)
420  54 13l 156 −22l
−13l −3l2 −22l 4l2
p̄0 l 
pTe = l
− 6l

1 6 1 (5.32)
2

It is easily verified that the stiffness matrix has one translational rigid body mode:

uT(1) = 1
 
0 1 0 (5.33)

and one rotational rigid body mode:

uT(2) = 1 − 2l l
 
1 2
(5.34)

The quadratic forms uT(1) Mu(1) and uT(2) Mu(2) are then equal to the translation and rotatory inertias

ml3
ml and 12 .

University of Liège -44-


Theory of Vibration 6. Solution methods for the eigenvalue problem

6 Solution methods for the eigenvalue problem


6.1 Describe the concept of reduction of the size of an eigenvalue problem
using the static condesation method.29

We start from the general matrix expression governing an eigenvalue problem:

Kx = ω 2 M x (6.1)

To reduce the size of K and M, we eliminate a subset of DOFs, which gives the nR remaining and nC

condensed coordinates respectively written xR and xC (nR  n). Eq. 6.1 is therefore partitionned as

follows:
     
KRR KRC xR 2 MRR MRC xR
=ω (6.2)
KCR KCC xC MCR MCC xC

Considering the inertia forces F = ω 2 M x, Eq. 6.2 becomes:


    
KRR KRC xR FR
= FC ' 0 (6.3)
KCR KCC xC FC

The inertia forces FC are negligible if the masses affected to the condensed DOF are negligible or equal

to zero, which gives:


−1
xC = −KCC KCR xR (6.4)

From there we can create the transformation matrix R:


     
xR xR I
x= = −1 = R xR = −1 xR (6.5)
xC −KCC KCR xR −KCC KCR

Finally, the general stiffness and mass matrices are given by:
(
−1
K = RT KR = KRR − KRC KCC KCR
T −1 −1 −1 −1
M = R M R = MRR − MRC KCC KCR − KRC KCC MCR + KRC KCC MCC KCC KCR

So we have the reduced eigenvalue problem:

(K − ω 2 M )xR = 0 (6.6)

Notice that the Guyan’s reduction is only valid if the inertial forces FC are indeed negligible, and that

it can be shown that this method always lead to an excess approximation of the eigenvalue spectrum

anyway.

In computational practice, the reduction matrix


   
I I
R= −1 = (6.7)
−KCC KCR RCR

is computed by solving the static problem (with nR second members): KCC RCR = −KCR .
29 Ref. book p.481-482 & slides 6.10-13

University of Liège -45-


Theory of Vibration 6. Solution methods for the eigenvalue problem

6.2 Describe the concept of mechanical impedance and how it can be used
for substructuring a finite element model.30

Large number of DOFs ⇒ dynamic analysis unfeasible. Hence we break down the system in several

substructures (Fig. 6.1) and we study their response to forces applied on their interface boundaries.

The result is that we have a reduced set of variable which approximate each substructure and by doing

so, the entire system.

Figure 6.1: Mechanical impedance concept

The impedance matrix yields:

Z(ω 2 ) = K − ω 2 M = H −1 (ω 2 ) (6.8)

and is the matrix which relates the applied force amplitudes s to the displacement amplitudes q as

follows:

s = Z(ω 2 )q (6.9)

Since the internal degrees of freedom q1 (cf. Fig. 6.1) are not externally loaded, we may write:

Z11 (ω 2 ) Z12 (ω 2 ) q1
    
0
= (6.10)
Z21 (ω 2 ) Z22 (ω 2 ) q2 s2

Taking the first equation of this system to eliminate the internal degrees of freedom gives:

−1
q1 = −Z11 Z12 q2 , (6.11)

which we can inject into the second equation to finally obtain:


Z22 (ω 2 )q2 = s2 (6.12)

where
∗ −1
Z22 = Z22 − Z21 Z11 Z12 . (6.13)


Notice that Z22 admits as poles the zeros of Z11 which are the eigenfrequencies of the subsystem with

its boundary degrees of freedom q2 fixed.


30 Ref. book p.105 & slides 2(part3).10-15

University of Liège -46-


Theory of Vibration 6. Solution methods for the eigenvalue problem

Considering the subsystem clamped on its boundary, we can compute the eigensolutions of the

subsystem:

(K11 − ω 2 M11 )x = 0 (6.14)

−1
Based on the spectral expansion of Z11 , it can be shown that the reduced impedance matrix takes the

form:
∗ −1
Z22 = K22 − K21 K11 K12

−1 −1 −1 −1
− ω 2 (M22 − M21 K11 K12 − K21 K11 M12 + K21 K11 M11 K11 K12 )
n1
X (K21 − ω̄s2 M21 ) x̄(s) x̄T(s) (K21 − ω̄s2 M21 )T
− ω4
s=1
ω̄s4 (ω̄s2 − ω̄ 2 )

Then one can deduce that the dynamics of the substructure may be described in terms of static modes

associated to the interface (support) displacements and in terms of the internal modes (when the

interface is fixed).

This forms the main idea underlying the Craig-Bampton method (substructuring a finite element

model).

University of Liège -47-


Theory of Vibration 7. Direct time-integration methods

7 Direct time-integration methods


7.1 Introduce the general formula for direct integration of first-order
systems. Define the concept of an explicit or implicit integration
scheme. 31

With the assumption of viscous dissipation in the Lagrange equations of the system:

1 T
D= q̇ C q̇ ≥0 (7.1)
2

the general dynamic equilibrium equation of the system writes:

M q̈ + C q̇ + K q = p(t) (7.2)

Direct multistep integration methods for first-order differential systems in the form:

ẏ = f (y, t) (7.3)

can be stated as:


m
X m
X
yn+1 = αj yn+1−j − h βj ẏn+1−j . (7.4)
j=1 j=0

where h = tn+1 − tn is the time step and yn+1 is the solution to equation (7.3) at time tn+1 calculated

from the solutions at the m preceding times, from their derivatives and from the derivative of yn + 1

itself.

For β0 6= 0, the time integration scheme is said to be implicit, since the solution at time tn+1 is a

function of its own time derivative. Therefore, the integration relationships have to be recast before

they can be solved. The solution method becomes iterative in the nonlinear case.

For β0 = 0, the time integration scheme is said to be explicit, since yn+1 can be deduced directly from

the results at the previous time step.

When αj = 0 and βj = 0 for j > 1, the integration scheme corresponds to a one-step method, since

the system at time tn+1 is only a function of its previous state at time tn .

7.2 Newmark
7.2.1 Applying Newmark’s integration formulas (equation(7.5)), describe the Newmark’s
integration algorithm for linear systems.32
q̇n+1 = q̇n + (1 − γ) h q̈n + γ h q̈n+1
(7.5)
2 2
qn+1 = qn + h q̇n + (0.5 − β) h q̈n + h β q̈n+1
Time-integration of the dynamic equilibrium equations yields:

M q̈ + C q̇ + Kq = p(t) (7.6)
31 Ch7,part1, slide 3-4 and Ref. book p.513-514
32 Ch7,part1, slide 17 and Ref. book p.522-523

University of Liège -48-


Theory of Vibration 7. Direct time-integration methods

At time tn+1 , and associated with the iteration matrix [M + γhC + βh2 K], we have the linear system

of equations:
   
2 1 2
[M + γhC + βh K]q̈n+1 = pn+1 − C(q̇n + (1 − γ)hq̈n ) − K qn + hq̇n + − β h q̈n (7.7)
2

7.2.2 According to Fig. 7.1 and the expressions of amplitude and periodicity error,
discuss the stability and accuracy of the method with respect to the choice of
parameters and time-step.33

Figure 7.1: Newmark algorithm

1
Fig. 7.1 shows that γ must be ≥ 2 at all time for the algorithm to be stable. Unconditionnal stability is

met if β ≥ 1/4 (1/2 + γ)2 as shown in Fig. 7.1. Conditionnal stability impose a time-step:

4
(γ + 0.5)2 − 4β ≤ 2
ωmax h2
Stability limit
⇔ h≤
ωmax

You can land on the values displayed in Tab. 7.1 just by injecting the values of the free parameters into

the equation.

- The purely explicit scheme is not used in practice, since it is in any case unstable.

- The Fox and Goodwin algorithm does not generate damping and leads to a periodicity error of fourth

order, but is conditionally stable.

33 Ch7, part1, slide 24-26 and Ref. book p.533-534,537-539

University of Liège -49-


Theory of Vibration 7. Direct time-integration methods

Table 7.1: Algorithm table

- The average constant acceleration algorithm is the best unconditionally stable scheme since it doesn’t

generate numerical damping and is characterized by a periodicity error of second order.

- The average constant acceleration scheme modified to introduce numerical damping is characterized

by a numerical damping ratio growing linearly with ωh. It is interesting to note that the first-order

error that results from taking γ 6= 1/2 manifests itself in the form of excessive numerical dissipation,

but not in period disrepancy since the latter remains of second-order for α 6= 0.

1 1
Finally, taking β = 6 or = 4 convey a better stability but increase the periodicity error as much as β

increase. The amplitude error will be non-existant as long as γ = 21 , which is convenient because it is

the value which gives the larger domain of unconditional stability. If γ 6= 12 , the amplitude error is given

by the numerical damping formula (cf. Fig. 7.1).

7.3 Describe the α-method of Hilber-Hughes-Taylor (HHT) and compare it


to Newmark’s method using Figure 7.2.3435
7.3.1 Description of the Hilber-Hughes-Taylor (a.k.a alpha-) method

It is used to allow the damping in the modelling without degrading the order of accuracy. The HHT-α

method is a generalisation of the Newmark method which corresponds to the case where the lag in the

damping is not equal to α and is null. Its principle consists in applying the Newmark’s formula (7.5)

whereas the time-discrete equilibrium equations are modified by averaging the internal (i.e. elastic and

damping) forces and the external forces between both time instants.

34 Ch7, part1, slides 27-29 and Ref. book p.539, 547-549


35 Because of the lack of information on the relative slides and since the book describes a different figure this answer
should be considered carefully.

University of Liège -50-


Theory of Vibration 7. Direct time-integration methods

With

M q̈ + C q̇ + K q = g(q,t) where f (q, q̇) = C q̇ + K q (7.8)

The adapted Newmark’s formula in a more general form is:

Mq̈n+1 + (1 − α)f (qn+1 , q̇n+1 ) + αf (qn , q̇n ) = (1 − α)g(qn+1 , tn+1 ) + αg(qn , tn ) (7.9)

The HHT-α is at least 2nd-order accurate and unconditionally stable if the following conditions are

satisfied:

 
1 1 1
γ= + α, β= (1 + α)2 , α∈ 0, (7.10)
2 4 3

The HHT-α method is useful in structural dynamics simulations incorporating many degrees of

freedom, and in which it is desirable to numerically attenuate (or dampen-out) the response at high

frequencies. Numerical damping (impossible in Newmark) stabilizes the numerical integration scheme

by damping out the unwanted high frequency modes and numerical noise.(36 )

7.3.2 Comparison

The Figure 7.2 shows the evolution of the numerical damping ratio through a frequency range i.e. how

each spectral component will be affected by the time integrator (=dispersion). One sees that the same

value of alpha and at a fixed frequency, a smaller timestep is needed in the Newmark algorithm for the

same numerical damping ratio. Indeed, the Newmark regular method introduces a numerical damping

ratio growing linearly with ωh while the curve computed for the HHT method appears to be quadratic.

Figure 7.2: Numerical damping ratio for Newmark and HHT

36 http://people.duke.edu/~hpgavin/cee541/NumericalIntegration.pdf#equation.0.31

University of Liège -51-


Theory of Vibration 7. Direct time-integration methods

7.4 Explain how Newmark’s algorithm can be extended to nonlinear


systems of general form as in Eq. (7.11) using, as a reminder,
Newmark’s integration formulas Eq. (7.12).37

Mq̈ + f (q, q̇) = p(q,t)
(7.11)
q0 , q˙0 given


q̇n+1 = q̇n + (1 − γ)h q̈n + γh q̈n+1
(7.12)
qn+1 = qn + h q̇n + h2 ( 21 − β) q̈n + h2 β q̈n+1

First, we define the residual vector r(q) through the equilibrium equations as

r(q) = Mq̈ + f (q, q̇) − p(q,t) = 0 (7.13)

∗ ∗
Since q̇n+1 = q̇n + (1 − γ)h q̈n and qn+1 = qn + h q̇n + h2 ( 21 − β) q̈n , the Newmark’s formula can be

rewritten as follows:
1


 q̈n+1 =
 (qn+1 − qn+1 )
βh2 (7.14)
∗ γ ∗
 q̇n+1 = q̇n+1
 + (qn+1 − qn+1 )
βh
k
The residual equation is expressed in terms of qn+1 only so r(qn+1 ) = 0. Let us denote qn+1 an

approximate value of qn+1 resulting from iteration k. In the neighbourhood of this value, the residual

equation can be linearized:

That leads to the iteration matrix given by

γ t 1
S(q) = Kt + C + M (7.15)
βh βh2

The solution of r(qn+1 ) = 0 can be found iteratively as explained in Fig. 7.3. At iteration k of the time

k
step n+1, we use r(qn+1 ) = 0 to compute ∆qk and its first and second derivatives:

γ 1
S∆qk = −r(qn+1
k
), ∆q̇k = ∆qk , ∆q̈k = ∆qk . (7.16)
βh βh2

k+1
From those variables, we can compute qn+1 and continue the loop.

37 Ch7, part2, slides 5-9 and Ref. book p.564-567

University of Liège -52-


Theory of Vibration 7. Direct time-integration methods

Figure 7.3: Implicit integration of the response (non linear case)

University of Liège -53-


Theory of Vibration 8. Introduction to nonlinear systems

8 Introduction to nonlinear systems


38
8.1 Describe the different sources of nonlinearity in mechanical systems.

• Material nonliearity

Two exemples of nonlinearity in materials are the nonlinear elastic materials (Such as rubber isolators),

or the dry friciton (It is applied to brakes, rotor-stator contact, drill-string systems,...).

Figure 8.1: Examples of material non-linearity

When we linearise, we obtain the slope at the point considered. We have to start again for each specific

point.

For small loads, we have a linear behavior. When non-linearity appear when we increase too much the

load since the frequency depends on the excitation.

• Geometric nonliearity, related to high displacement

A common example of non-linearity is the simple pendulum. In case of large amplitudes, there are

changes in the frequency of oscillation and apparition of harmonic distortions.

Figure 8.2: Simple pendulum.

If we compute for different initial conditions, we will have different results. When we increase the initial

angle, the fundamental frequency gets lower accordingly. We can also see the apparition of resonance

frequencies and harmonics.


38 Chapter 8, slides 4-23

University of Liège -54-


Theory of Vibration 8. Introduction to nonlinear systems

• Nonliearity at boundary conditions

We can linearize for a given load only since the load varies with the variation of stiffness.

Figure 8.3: Clearance and contact.

• Nonliear body forces

Figure 8.4: Parallel plate capacitor with gap d.

• Nonliear excitation

39
8.2 Explain what the main challenges are in nonlinear dynamics.

The study of nonlinear vibrating structures is not mature.

1. Mathematical theory is made difficult

• No principle of superposition

• Non-unicity of the solution

2. Rich, complex, ’exotic’ behaviour

• Frequency-energy dependence

• Jump phenomena → bifurcations

• Harmonics

• Chaos

→ Interpretation and use for design of engineering structures are still a challenge.

39 Chapter 8, slides 24-28

University of Liège -55-


Theory of Vibration 8. Introduction to nonlinear systems

8.2.1 No principle of superposition

Linear system → If it obeys principle of superposition. This means:

1. Output directly proportional to the input of the system

2. If input signal = summation of several different waveforms → Output = summation of individual

output responses

Figure 8.5: Single DOF Linear / Non-linear comparison

The non linear model becomes frequency-energy dependent.

Fig. 8.6, clearly shows the effect of the cubic non-linearity in the free response of the mass:

Figure 8.6: Single DOF Linear / Non-linear free response comparison (Left->Linear Right->Non-linear)

University of Liège -56-


Theory of Vibration 8. Introduction to nonlinear systems

8.3 Apply the harmonic balance method to the example of the Duffing
oscillator to calculate an approximation of its free response. 40

Figure 8.7: Equations of the Duffing Oscillator with initial conditions

We start by considering the approximation of the free response in the form:

x(t) = A cos(ωt) (8.1)

Introducing it into the equation of motion gives:

− ω 2 m A cos(ωt) + k A cos(ωt) + knl A3 cos3 (ωt) = 0 (8.2)

Knowing that:
3
eiωt + e−iωt

3 1
cos3 (ωt) = = cos(ωt) + cos(3ωt) (8.3)
2 4 4

And introducing it into Eq. 8.2, one obtains:


 
2 3 3 1
− ω m A cos(ωt) + k A cos(ωt) + knl A cos(ωt) + cos(3ωt) = 0 (8.4)
4 4

Balancing the terms in cos(ωt) (and neglecting the term in cos(3ωt):

 3
− ω 2 m + k + knl A2 A cos(ωt) = 0 (8.5)
4
3
− ω 2 m + k + knl A2 =0 (8.6)
4

We get the non-trivial solution:


r
4 
A≈ ω2 m − k (8.7)
3knl

Or, solving for ω:


r
k 3 knl 2
ω0 ≈ + A (8.8)
m 4 m

Equation (8.8) shows that the natural frequency of the nonlinear system depends on the amplitude of

the oscillation. With equation (8.8), it is also possible to recover to the typical result in two specific

cases:
q
k
• When knl tends to zero: if knl → 0 then ω0 → m
q
k
• For small displacements: if A → 0 then ω0 → m

40 Chapter 8 slides 29-31

University of Liège -57-

You might also like