You are on page 1of 22

LECTURE NOTES

EMT 2502 VIBRATIONS


MECHATRONIC ENGINEERING
Y5 S1
DeKUT

By
Inno Odira


c Copyright by Inno Odira, 2018
Table of Contents

Table of Contents iii

1 Multi DegreeDegree of Freedom Vibrating Systems (MDOF) 1


1.1 Energy Methods of Modelling . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.2 Lagrange’s Equation (Without external forces nor damping) . . . . . 2
1.1.3 Lagrange’s Method (With Viscous damping but Without external forces
) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1.4 Lagrange’s Method (With external forces ) . . . . . . . . . . . . . . . 3
1.2 Influence Coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

Bibliography 20

REFERENCES 20

iii
Chapter 1

Multi DegreeDegree of Freedom


Vibrating Systems (MDOF)

1.1 Energy Methods of Modelling


1.1.1 Introduction
• Energy methods allow us to ignore internal/ interaction forces

• Energy methods are based on:

– Generalised coordinates: Minimal set of independent coordinates needed to


describe the system
– Virtual displacements: Used in derivation and for accounting for external forces

Generalised Coordinates

Figure 1.1: Illustration of generalised coordinates

Consider Fig. 1.1.


We can choose (X, θ) or (X1 , X2 ) as generalised coordinates.

1
2
Generalised coordinates are often written as

q̄ = q1 , q2 , · · · , qn

For our example


q̄ = (X, θ) or q̄ = (X1 , X2 )

1.1.2 Lagrange’s Equation (Without external forces nor damping)


Define The Lagrangian

L = T − V = Kinetic Energy − Potential Energy


Lagrange’s Equation becomes
 
d ∂L ∂L
− = 0, i = 1, 2, · · · , n (1.1)
dt ∂ q̇i ∂qi

Where n is the number of generalised coordinates (Number of DOF)


Note: Be Sure to include all system energies and define them consistently

Example:
Consider a simple mass- spring system Fig. 1.2

Figure 1.2: Mass-Spring system

The generalised coordinates


q̄ = q1 = x
Form the Kinetic andPotential Energies in terms of the generalised coordinates

1 1
T = mẋ2 V = kx2
2 2
3
Lagrangian

1 1
L = mẋ2 − kx2
2 2
Plug Lagrangian in Lagrange’s Equation

 
d ∂L ∂L
− = 0
dt ∂ q̇i ∂qi
 
d ∂L ∂L
− = 0
dt ∂ ẋi ∂xi
d
(mẋ) − kx = 0
dt
mẍ − kx = 0

This matches what we get with Newton/ Euler Method

1.1.3 Lagrange’s Method (With Viscous damping but Without


external forces )
Remember damping force (FD )
FD = cδ̇
Where δ̇ is the rate of change of the length of the damper.
To match the form needed for lagrange, integrate with respect to δ̇
Z Z
1
FD dδ̇ = (cδ̇)dδ̇ = cδ̇ 2
2
This gives Rayleigh’s Dissipation Function (RD)- (Only valid for linear viscous damping)

1
RD = cδ̇ 2
2
Use this to represent the damper in Lagrange’s Equation ie.
 
d ∂L ∂RD ∂L
+ − = 0, i = 1, 2, · · · , n (1.2)
dt ∂ q̇i ∂ q̇i ∂qi

1.1.4 Lagrange’s Method (With external forces )


To understand external forces, we need to introduce virtual displacements
4
Virtual Displacement

• Infinitesimally small changes in generalised coordinates

• Occur in zero time(No time elapses during the move)

• Do not violate system constraints

Consider f (q1 , q2 , · · · , t) = c as a system constraint. If we allow virtual displacements


δq1 , δq2 , · · · then f (q1 + δq1 , q2 + δq2 , · · · , t
|{z} )= c
|{z}
No Time has passed Constraint still holds
This gives us a way to determine Virtual work which helps us on how to determine the
contribution of each external force in the direction of each generalised coordinates, ie.

δWi = Qi δqi
Where Qi = Component of the external force acting in the direction δqi
Qi goes into the Lagrange’s Equation
We can write the total form of Lagrange’s Equation
 
d ∂L ∂RD ∂L
+ − = Qi , i = 1, 2, · · · , n (1.3)
dt ∂ q̇i ∂ q̇i ∂qi

Example
Consider a 3 DOF mass- spring-Damper system Fig. 1.3

Figure 1.3: 3 DOF Mass-Spring -Damper system

Example

The system has 3 Bodies, implies 3 DOF and 3 generalised coordinates and 3 Equations
of Motion.
use generalised coordinates q = x1 , x2 and x3
The kinetic energy of the system at an arbitrary instant is
1 1 1
T = mẋ21 + 2mẋ22 + mẋ23
2 2 2
5
Potential Energy of the system at an arbitrary instant is

1 1 1 1
V = kx21 + 2k(x2 − x1 )2 + k(x3 − x2 )2 + 3kx23
2 2 2 2
Lagrangian is

1 2
mẋ1 + 2mẋ22 + mẋ23 − kx21 − 2k(x2 − x1 )2 − k(x3 − x2 )2 − 3kx23

L=
2
Rayleigh’s dissipation function
1 1 1 1
RD = − cẋ21 − 2c(ẋ2 − ẋ1 )2 − c(ẋ3 − ẋ2 )2 + 3cẋ23
2 2 2 2
The work done by external forces is

δW = F1 (t)δx1 + F2 (t)δx2
Thus
Q1 = F1 (t), Q2 = F2 (t) and Q3 = 0
Application of Lagrange’s Equation for x1 leads to
 
d ∂L ∂RD ∂L
+ − = Q1
dt ∂ ẋi ∂ ẋi ∂xi
   
d 1 1 1
m(2)ẋ1 − − c(2)ẋ2 − 2c(2)(ẋ2 − ẋ1 )(−1)
dt 2 2 2
 
1 1
− − k(2)x2 − 2k(2)(x2 − x1 )(−1) = F1 (t) (1.4)
2 2

Application of Lagrange’s Equation for x2 leads to


 
d ∂L ∂RD ∂L
+ − = Q2
dt ∂ ẋi ∂ ẋi ∂xi
   
d 1 1 1
2m(2)ẋ2 − − 2c(2)(ẋ2 − ẋ1 ) − c(2)(ẋ3 − ẋ2 )(−1)
dt 2 2 2
 
1 1
− − 2k(2)(x2 − x1 ) − k(2)(x3 − x2 )(−1)) = 0 (1.5)
2 2

Application of Lagrange’s Equation for x3 leads to


 
d ∂L ∂RD ∂L
+ − =0
dt ∂ ẋi ∂ ẋi ∂xi
   
d 1 1 1
m(2)ẋ3 − − c(2)(ẋ3 − ẋ2 ) − 3c(2)ẋ3
dt 2 2 2
 
1 1
− − k(2)(x3 − x2 ) − 3k(2)x3 = F2 (t) (1.6)
2 2
6
Rearranging (1.4), (1.5) and (1.6) and summarizing in matrix form leads to

          
m 0 0 ẍ1 3c −2c 0 ẋ1 3k −2k 0 x1 F1 (t)
          
 0 2m 0   ẍ2 + −2c 3c −c   ẋ2 + −2k 3k −k   x2  =  F2 (t) 
          
0 0 m ẍ3 0 −c 4c ẋ3 0 −k 4k x3 0
(1.7)

Matrix Formulation

M ẍ + C ẋ + Kx = F (1.8)
Where
M is the n × n Mass matrix
K is the n × n Stiffness matrix
C is the n × n Damping matrix
F is the n × 1 Force vector
x is the n × 1 Displacement vector
·x is the n × 1 Velocity vector
ẍ is the n × 1 Acceleration vector

1.2 Influence Coefficients


Introduction
• The equations of motion of a multidegree-of-freedom system can also be written in
terms of influence coefficients, which are extensively used in structural engineering.
Basically, one set of influence coefficients can be associated with each of the matrices
involved in the equations of motion.

• The influence coefficients associated with the stiffness and mass matrices are, respec-
tively, known as the stiffness and inertia influence coefficients.

• In some cases, it is more convenient to rewrite the equations of motion using the
inverse of the stiffness matrix (known as the flexibility matrix) or the inverse of the
mass matrix.

• The influence coefficients corresponding to the inverse stiffness matrix are called the
flexibility influence coefficients,

• and those corresponding to the inverse mass matrix are known as the inverse inertia
coefficients.
7
Stiffness influence coefficients
• For a simple linear spring, the force necessary to cause a unit elongation is called the
stiff-ness of the spring.

• In more complex systems, we can express the relation between the dis-placement at a
point and the forces acting at various other points of the system by means of stiffness
influence coefficients.

• The stiffness influence coefficient, denoted as kij , is defined as the force at point i due
to a unit displacement at point j when all the points other than the point j are fixed.
Consider a Multidegree-of-freedom spring-mass system. Fig. 1.4

Figure 1.4: Multidegree-of-freedom spring-mass system.

Using the definition above, for the spring-mass system shown in Fig. 1.4, the total force at
point i, Fi can be found by summing up the forces due to all displacements xj (j = 1, 2, · · · , n)
as

Fi = Σnj=1 kij xj i = 1, 2, · · · , n (1.9)


Equation (1.9) can be stated in matrix form as


F = [k]→
−x (1.10)


where → −x and F are the displacement and force vectors defined in Eq. (6.7) and [k] is
the stiffness matrix given by
 
k11 k12 · · · k1n
 
 k21 k22 · · · k2n 
[k] =  (1.11)
 
.. 

 . 

kn1 kn2 · · · knn
8
The following aspects of stiffness influence coefficients are to be noted:

1. Since the force required at point i to cause a unit deflection at point j and zero deflection
at all other points is the same as the force required at point j to cause a unit deflection
at point i and zero deflection at all other points (Maxwell s reciprocity theorem [6.1]),
we have kij = kji .

2. The stiffness influence coefficients can be calculated by applying the principles of statics
and solid mechanics.

3. The stiffness influence coefficients for torsional systems can be defined in terms of
unit angular displacement and the torque that causes the angular displacement. For
example, in a multirotor torsional system, kij can be defined as the torque at point i
(rotor i) due to a unit angular displacement at point j and zero angular displacement
at all other points.

The stiffness influence coefficients of a multidegree-of-freedom system can be determined


as follows:

1. Assume a value of one for the displacement x( j = 1to start with) and a value of zero
for all other displacements x1 , x2 , · · · xj−1 , xj+1 , · · · , xn By definition, the set of forces
kij (i = 1, 2, · · · , n) will maintain the system in the assumed configuration (xj = 1, x1 =
x2 = · · · = xj − 1 = xj+1 = · · · = xn = 0) Then the static equilibrium equations are
written for each mass and the resulting set of n equations solved to find the n influence
coefficients kij (i = 1, 2, · · · , n).

2. After completing step 1 for j = 1,the procedure is repeated for j = 2, 3, · · · , n.

The following example illustrates the procedure

Example
Consider a Multidegree-of-freedom spring-mass system. Fig. 1.5 Find the stiffness influence
coefficients of the system shown in Fig. 1.5
Solution
Approach: Use the definition of kij and static equilibrium equations. Let x1 , x2 , and x3
denote the displacements of the masses m1 , m2 and m3 , respectively. The stiffness influence
coefficients kij of the system can be determined in terms of the spring stiffnesses k1 , k2 ,
and k3 as follows. First, we set the displacement of m1 equal to one (x1 = 1) and the
displacements of m2 and m3 equal to zero (x2 = x3 = 0), as shown in Fig. 1.6.
The set of forces ki1 (i = 1, 2, 3) is assumed to maintain the system in this configuration.
The free-body diagrams of the masses corresponding to the configuration of Fig.1.6. are
indicated in Fig. 1.7.
9

Figure 1.5: Determination of stiffness influence coefficients..

Figure 1.6: Determination of stiffness influence coefficients..

Figure 1.7: Determination of stiffness influence coefficients..

The equilibrium of forces for the masses m1 , m2 and m3 in the horizontal direction yields

Mass m1 : k1 = −k2 + k11 (1.12)

Mass m2 : k21 = −k2 (1.13)

Mass m3 : k31 = 0 (1.14)


The solution of Eqs. (1.12) to (1.14) gives

k11 = k1 + k2 , k21 = −k2 , k31 = 0 (1.15)


10
Next the displacements of the masses are assumed as x1 = 0, x2 = 1, and x3 = 1 as in
Fig.1.8

Figure 1.8: Determination of stiffness influence coefficients..

Since the forces ki2 (i = 1, 2, 3) are assumed to maintain the system in this configuration,
the free-body diagrams of the masses can be developed as indicated in Fig. 1.9.

Figure 1.9: Determination of stiffness influence coefficients..

The equilibrium of forces for the masses m1 , m2 and m3 in the horizontal direction yields

Mass m1 : k12 + k2 = 0 (1.16)

Mass m2 : k22 − k3 = k2 (1.17)

Mass m3 : k32 = −k3 (1.18)


The solution of Eqs. (1.16) to (1.18) gives

k12 = −k2 , k22 = k2 + k3 , k32 = −k3 (1.19)


Finally the set of forces ki3 (i = 1, 2, 3) is assumed to maintain the system with x1 =
0, x2 = 0, and x3 = 1 as in Fig. 1.10
The free-body diagrams of the various masses in this configuration are shown in Fig. 1.11
11

Figure 1.10: Determination of stiffness influence coefficients..

Figure 1.11: Determination of stiffness influence coefficients..

and the force equilibrium equations lead to

Mass m1 : k13 = 0 (1.20)

Mass m2 : k23 + k3 = 0 (1.21)

Mass m3 : k33 = k3 (1.22)


The solution of Eqs. (1.20) to (1.22) gives

k13 = 0, k23 = −k3 , k33 = k3 (1.23)


Thus the stiffness matrix of the system is given by
 
(k1 + k2 ) −k2 0
 
[k] = 
 −k2 (k2 + k3 ) −k3 
 (1.24)
0 −k3 k3

Flexibility influence coefficients


• As seen in Previous Examples the computation of stiffness influence coefficients requires
the application of the principles of statics and some algebraic manipulation. In fact, the
12
generation of n stiffness influence coefficients k1j , k2j , · · · , knj for any specific j requires
the solution of n simultaneous linear equations. Thus n sets of linear equations (n
equations in each set) are to be solved to generate all the stiffness influence coefficients
of an n-degree-of-freedom system. This implies a significant computational effort for
large values of n.

• The generation of the flexibility influence coefficients, on the other hand, proves to be
simpler and more convenient. To illustrate the concept, consider again the spring-mass
system shown in Fig. 1.4

Let the system be acted on by just one force Fj and let the displacement at point i (i.e.,
mass mi ) due to Fj be xij . The flexibility influence coefficient, denoted by aij is defined
as the deflection at point i due to a unit load at point j. Since the deflection increases
proportionately with the load for a linear system, we have

xij = aij Fj (1.25)


If several forces Fj (j = 1, 2, · · · , n) act at different points of the system, the total deflec-
tion at any point i can be found by summing up the contributions of all forces Fj :
n
X n
X
xi = xij = aij Fj , i = 1, 2, · · · , n (1.26)
j+1 j+1

Equation (1.26) can be expressed in matrix form as


− →

x = [a] F (1.27)


where → −
x and F are the displacement and force vectors defined in Eq. (6.7) and [a]
is the flexibility matrix given by
 
a11 a12 · · · a1n
 
 a21 a22 · · · a2n 
[a] =  (1.28)
 
.. 

 . 

an1 an2 · · · ann
The following characteristics of flexibility influence coefficients can be noted:

1. An examination of Eqs. (1.27) and (1.24) indicates that the flexibility and stiffness
matrices are related. If we substitute Eq. (1.27) into Eq. (1.24), we obtain


− →

x = [a] F = [a][k]→

x (1.29)

from which we can obtain the relation


13

[a][k] = [I] (1.30)

where [I] denotes the unit matrix. Equation ((1.30) is equivalent to

[k] = [a]−1 , [a] = [k]−1 (1.31)

That is, the stiffness and flexibility matrices are the inverse of one another

2. Since the deflection at point i due to a unit load at point j is the same as the deflection
at point j due to a unit load at point i for a linear system (Maxwell s reciprocity theo-
rem [6.1]), we have aij = aji .

3. The flexibility influence coefficients of a torsional system can be defined in terms of


unit torque and the angular deflection it causes. For example, in a multirotor torsional
system, aij can be defined as the angular deflection of point i (rotor i) due to a unit
torque at point j (rotor j).

The flexibility influence coefficients of a multidegree-of-freedom system can be determined


as follows:

1. Assume a unit load at point j(j = 1 to start with) By definition, the displacements of
the various points i(i = 1, 2, · · · , n) resulting from this load give the flexibility influence
coefficients, aij i = 1, 2, · · · , n. Thus aij can be found by applying the simple principles
of statics and solid mechanics.

2. After completing Step 1 for j = 1, the procedure is repeated for j = 2, 3, · · · , n.

3. Instead of applying Steps 1 and 2, the flexibility matrix, [a], can be determined by
finding the inverse of the stiffness matrix, [k], if the stiffness matrix is available.

The following example illustrates the procedure

Example
Consider a Multidegree-of-freedom spring-mass system. Fig. 1.12 Find the flexibility influ-
ence coefficients of the system shown in Fig. 1.12
Solution
Approach: Use the definition of kij and static equilibrium equations. Let x1 , x2 , and x3
denote the displacements of the masses m1 , m2 and m3 , respectively. The Flexibility
influence coefficients aij of the system can be determined in terms of the spring stiffnesses
k1 , k2 , and k3 as follows. Apply a unit force at mass m1 and no force at other masses
(F1 = F2 = F3 = 0) equal to zero (x2 = x3 = 0), as shown in Fig. 1.13.
14

Figure 1.12: Determination of flexibility influence coefficients.

Figure 1.13: Determination of flexibility influence coefficients.

The resulting deflections of the masses m1 , m2 , and m3 (x1 , x2 , and, x3 ) are, by definition,
a11 , a21 , and a31 , respectively (see Fig. 1.14). The free-body diagrams of the masses are
shown in Fig. 1.15 .

Figure 1.14: Determination of flexibility influence coefficients.

Figure 1.15: Determination of flexibility influence coefficients.


15
The equilibrium of forces in the horizontal direction for the various masses gives the
following:

Mass m1 : k1 a11 = k2 (a21 − a11 + 1) (1.32)

Mass m2 : k2 (a21 − a11 ) = k3 (a31 − a21 ) (1.33)

Mass m3 : k3 (a31 − a21 ) = 0 (1.34)


The solution of Eqs. (1.32) to (1.34) gives

1 1 1
a11 = , a21 = , a31 = (1.35)
k1 k1 k1
Next, we apply a unit force at mass m2 and no force at masses m1 and m3 as shown
in Fig. 1.16. These forces cause the masses m1 , m2 and m3 to deflect by x1 = a12 , x2 =
a22 , and x3 = a32 ,respectively (by definition of ai2 as shown in Fig. 1.16. The result-
ing free-body diagrams of the various masses as shown in Fig. (1.17) yield the following
equilibrium equations:

Figure 1.16: Determination of flexibility influence coefficients.

Figure 1.17: Determination of flexibility influence coefficients.

Mass m1 : k1 a12 = k2 (a22 − a12 ) (1.36)


16

Mass m2 : k2 (a22 − a12 ) = k3 (a32 − a22 + 1) (1.37)

Mass m3 : k3 (a32 − a22 ) = 0 (1.38)


The solution of Eqs. (1.40) to (1.42) gives

1 1 1 1 1
a12 = , a22 = + , a32 = + (1.39)
k1 k1 k2 k1 k2
Finally, when we apply a unit force to mass m3 and no force to masses m1 and m2
the masses deflect by x1 = a13 , x2 = a23 , and x3 = a33 , as shown in Fig. 1.18. The
resulting free-body diagrams of the various masses (Fig. 1.19) yield the following equilibrium
equations:

Figure 1.18: Determination of flexibility influence coefficients.

Figure 1.19: Determination of flexibility influence coefficients.

Mass m1 : k1 a113 = k2 (a23 − a13 ) (1.40)

Mass m2 : k2 (a23 − a13 ) = k3 (a33 − a23 ) (1.41)

Mass m3 : k3 (a33 − a23 ) = 1 (1.42)


The solution of Eqs. (1.40) to (1.42) gives
17
1 1 1 1 1 1
a13 = , a23 = + , a33 = + (1.43)
k1 k1 k2 k1 k2 k3
It can be verified that the stiffness matrix of the system, given by Eq. (E.13) of Example
6.3, can also be found from the relation [k] = [a] - 1.

Inertial influence coefficients


• The elements of the mass matrix, mij , are known as the inertia influence coefficients.
Although it is more convenient to derive the inertia influence coefficients from the
expression for kinetic energy of the system,the coefficients mij , can be computed using
the impulse-momentum relations.

• The inertia influence coefficients m1j , m2j , · · · , mnj are defined as the set of impulses
applied at points i − 1, 2, · · · , n respectively, to produce a unit velocity at point j and
zero velocity at every other point (that is, j = 1, ẋ1 = ẋ2 = · · · = ẋj−1 = ẋj+1 = · · · =
xn = 0 Thus, for a multidegree-of-freedom system, the total impulse at point i , Fi can
be found by summing up the impulses causing the velocities ẋj (j = 1, 2, · · · , n) as

n
X
Fi = mij ẋj (1.44)
j−1

Equation (1.44) can be stated in matrix form as



F = [m]→
−̇
x (1.45)


where →
−̇
x and F are the velocity and impulse vectors given by



F = (1.46)

   
ẋ1 Ḟ1
   

−  ẋ2  →
−  Ḟ2 
ẋ =  , F = (1.47)
   
.. .. 

 . 


 . 

ẋn Ḟn
and [m] is the mass matrix given by
 
m11 m12 · · · m1n
 
 21 m22 · · ·
 m m2n 
[m] =  . (1.48)

 .. .. .. 
 . ··· . 

mn1 mn2 · · · mnn
18
It can be verified easily that the inertia influence coefficients are symmetric for a linear
system that is, mij = mji . The following procedure can be used to derive the inertia influence
coefficients of a multidegree-of-freedom system.

1. Assume that a set of impulses fij are applied at various points i(i = 1, 2, · · · , n)
so as to produce a unit velocity at pointj(ẋj = 1with j = 1to start with and a
zero velocity at all other points (ẋ1 = ẋ2 = · · · ẋj−1 = ẋj+1 = · · · = ẋn = 0). By
definiion, the set of impulses fij (i = 1, 2, · · · , n) denote the inertia influence coefficients
mij (i = 1, 2, · · · , n).

2. After completing step 1 for j = 1, the procedure is repeated for j = 2, 3, · · · , n.

Note that if xj denotes an angular coordinate, then ẋj represents an angular velocity
and Fj indicates an angular impulse. The following example illustrates the procedure of
generating mij . Example
Consider a Compound pendulum and trailer system. Fig. 1.20 Find the inertia influence
coefficients of the system shown in Fig. 1.20

Figure 1.20: Compound pendulum and trailer system.

Solution
Approach: Use the definition of mij along with impulse-momentum relations. Let x(t) and
θ(t) denote the coordinates to define the linear and angular positions of the trailer (M )
and the compound pendulum (m). To derive the inertia influence coefficients, impulses
19

Figure 1.21: Compound pendulum and trailer system.

of magnitudes m11 and m21 are applied along the directions x(t) and θ(t) to result in the
velocities ẋ = 1 and θ̇ = 0 Then the linear impulse-linear momentum equation gives

m11 = (M + m)(1) (1.49)


and the angular impulse-angular momentum equation (about O) yields

l
m21 = (m)(1) (1.50)
2
and the angular impulse-angular momentum equation (about O) gives

ml2
m22 = (1) (1.51)
3
Thus the mass or inertia matrix of the system is given by
" #
M + m ml 2
[m] = 2
(1.52)
ml ml
2 2
REFERENCES

20

You might also like