0% found this document useful (0 votes)
13 views

Lecture Notes in Computational Mechanics

The lecture notes cover various topics in computational mechanics, including the stiffness method, finite element formulation of elastic continuum, and the development of truss and beam members. Key concepts such as element stiffness matrices, boundary conditions, and potential energy approaches are discussed in detail. The document serves as a comprehensive guide for understanding the principles and applications of computational mechanics.

Uploaded by

rupakhety.rajesh
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
13 views

Lecture Notes in Computational Mechanics

The lecture notes cover various topics in computational mechanics, including the stiffness method, finite element formulation of elastic continuum, and the development of truss and beam members. Key concepts such as element stiffness matrices, boundary conditions, and potential energy approaches are discussed in detail. The document serves as a comprehensive guide for understanding the principles and applications of computational mechanics.

Uploaded by

rupakhety.rajesh
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 106

Computational Mechanics 1

Lecture notes

Rajesh Rupakhety & Ragnar Sigbjörnsson


Contents

1 Introduction to the Stiffness method 4


1.1 Linear spring element . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.1.1 Element stiffness matrix . . . . . . . . . . . . . . . . . . . . . . . . 5
1.1.2 Boundary condition . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2 An assembly of linear springs . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2.1 Element stiffness matrices . . . . . . . . . . . . . . . . . . . . . . . 7
1.2.2 Assembly of system equations . . . . . . . . . . . . . . . . . . . . . 8
1.2.3 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2.4 Solution of unknown displacements . . . . . . . . . . . . . . . . . . 9
1.2.5 Solution of support reactions . . . . . . . . . . . . . . . . . . . . . . 9
1.3 Stepped bar: a discretized continuous system . . . . . . . . . . . . . . . . . 10
1.3.1 Discretization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.3.2 Formulation of element equations . . . . . . . . . . . . . . . . . . . 10
1.3.3 Assembly of system equations . . . . . . . . . . . . . . . . . . . . . 11
1.3.4 Boundary condition . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3.5 Solution of unknown displacements . . . . . . . . . . . . . . . . . . 11
1.3.6 Solution of unknown support reactions . . . . . . . . . . . . . . . . 12
1.3.7 Finding the displacement field . . . . . . . . . . . . . . . . . . . . . 12
1.3.8 Finding the strain field . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.3.9 Finding the stress field . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.4 Tapering bar: a continuous example . . . . . . . . . . . . . . . . . . . . . . 12
1.4.1 Discretization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.4.2 Formulation of element equations, assembly of system equations,
and solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.5 Potential energy approach . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.5.1 Principle of stationary potential energy . . . . . . . . . . . . . . . . 15
1.5.2 Stiffness matrix of a spring . . . . . . . . . . . . . . . . . . . . . . . 16
1.5.3 Assembly of linear springs . . . . . . . . . . . . . . . . . . . . . . . 17
1.5.4 Example of a truss structure . . . . . . . . . . . . . . . . . . . . . . 18
1.5.4.1 Solution by applying minimum potential energy principle
to the whole structure . . . . . . . . . . . . . . . . . . . . 19
1.5.4.2 Solution by assembly of element stiffness matrices . . . . . 22
1.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

2 Finite element formulation of elastic continuum 27


2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2 Strain tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.3 Stress tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

1
CONTENTS 2

2.4 Constitutive relation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30


2.5 Elastic strain energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.6 Work done by external forces . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.7 Potential energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.8 Formulation of element equations . . . . . . . . . . . . . . . . . . . . . . . 33
2.9 Formulation of system equations . . . . . . . . . . . . . . . . . . . . . . . . 35
2.10 Transformation of coordinate system . . . . . . . . . . . . . . . . . . . . . 36
2.11 Application of displacement boundary conditions . . . . . . . . . . . . . . 36
2.12 Computing the strain and stress field . . . . . . . . . . . . . . . . . . . . . 38
2.13 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

3 Development of truss members 39


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.2 Truss element . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.3 Shape function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.4 Stiffness matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.5 Plane truss : coordinate transformation . . . . . . . . . . . . . . . . . . . . 41
3.6 Space truss : coordinate transformation . . . . . . . . . . . . . . . . . . . . 42
3.6.1 Solved example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

4 Development of beam members 50


4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.2 Beam element . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.2.1 Mathematical model: Bernoulli-Euler beam . . . . . . . . . . . . . 51
4.2.1.1 Element coordinate system . . . . . . . . . . . . . . . . . 51
4.2.1.2 Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.2.1.3 Shape function . . . . . . . . . . . . . . . . . . . . . . . . 53
4.2.1.4 Strain matrix . . . . . . . . . . . . . . . . . . . . . . . . . 55
4.2.1.5 Elasticity matrix . . . . . . . . . . . . . . . . . . . . . . . 55
4.2.1.6 Element stiffness matrix . . . . . . . . . . . . . . . . . . . 55
4.2.1.7 Consistent force vector . . . . . . . . . . . . . . . . . . . . 56

5 FE formulation of dynamic systems 58


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.2 Terminology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.3 Kinetic energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.4 Lagrangian functional . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.5 Hamilton’s principle: Element equations . . . . . . . . . . . . . . . . . . . 60
5.6 Coordinate transformation and system equations . . . . . . . . . . . . . . . 61
5.7 Mass matrix of some common structural members . . . . . . . . . . . . . . 62
5.7.1 Truss element . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.7.2 Plane beam element . . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.7.3 Plane frame element . . . . . . . . . . . . . . . . . . . . . . . . . . 64
5.7.4 Space frame element . . . . . . . . . . . . . . . . . . . . . . . . . . 64

6 Free vibration of SDOF system 65


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
6.2 Free vibration of SDOF systems . . . . . . . . . . . . . . . . . . . . . . . . 65
6.2.1 System configuration . . . . . . . . . . . . . . . . . . . . . . . . . . 65
CONTENTS 3

6.2.2 Free vibration equation and solution . . . . . . . . . . . . . . . . . 66


6.3 Damped free vibration of SDOF systems . . . . . . . . . . . . . . . . . . . 68
6.3.1 System configuration . . . . . . . . . . . . . . . . . . . . . . . . . . 68
6.3.2 Damped free vibration equation and solution . . . . . . . . . . . . . 69
6.3.2.1 Case 1 : Critically damped system . . . . . . . . . . . . . 69
6.3.2.2 Case 2 : Overdamped system . . . . . . . . . . . . . . . . 72
6.3.2.3 Case 3: Underdamped system . . . . . . . . . . . . . . . . 74
6.3.3 Effect of damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
6.3.4 Measurement of damping . . . . . . . . . . . . . . . . . . . . . . . . 77

7 Forced vibration of SDOF systems 80


7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
7.2 Response to harmonic excitations . . . . . . . . . . . . . . . . . . . . . . . 80
7.2.1 Harmonic response of undamped systems . . . . . . . . . . . . . . . 80
7.2.1.1 Steady state response and displacement response factor . . 82
7.2.1.2 Resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
7.2.2 Harmonic response of damped systems . . . . . . . . . . . . . . . . 84
7.2.2.1 Steady state response and displacement response factor . . 87
7.2.2.2 Resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

8 Free vibration of MDOF systems 91


8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
8.2 Vibration frequencies and mode shapes . . . . . . . . . . . . . . . . . . . . 91
8.2.1 Reduction to a standard eigenvalue problem . . . . . . . . . . . . . 92
8.3 Spectral and mode shape matrices . . . . . . . . . . . . . . . . . . . . . . . 92
8.4 Static condensation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
8.5 Orthogonality of mode shapes . . . . . . . . . . . . . . . . . . . . . . . . . 94
8.6 Generalized mass and generalized stiffness matrix . . . . . . . . . . . . . . 96
8.7 Normalization of mode shapes . . . . . . . . . . . . . . . . . . . . . . . . . 96
8.8 Free vibration solution by modal expansion . . . . . . . . . . . . . . . . . . 97
8.8.1 Modal expansion of displacement . . . . . . . . . . . . . . . . . . . 98
8.8.2 Modal decoupling of the equation of motion . . . . . . . . . . . . . 99
8.9 Damping models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
8.9.1 Classically damped systems . . . . . . . . . . . . . . . . . . . . . . 100
8.9.1.1 Mass proportional damping . . . . . . . . . . . . . . . . . 100
8.9.1.2 Stiffness proportional damping . . . . . . . . . . . . . . . 101
8.9.1.3 Rayleigh damping . . . . . . . . . . . . . . . . . . . . . . 102
8.9.1.4 Damping matrix by superposition of modal damping ratios 103
8.9.2 Non-classically damped systems . . . . . . . . . . . . . . . . . . . . 104
8.10 Free vibration of classically damped MDOF systems . . . . . . . . . . . . . 104
Chapter 1

Introduction to the Stiffness method

Introduction
This chapter introduces the basic concepts of the direct stiffness method used in solving
structural mechanics problems. The linear spring is introduced first as it provides a
simple yet instructive example of the basic concepts. Some basic terminology such as
stiffness, degrees of freedom, stiffness matrix, etc are presented. The formation of stiffness
matrix of a linear spring element from the force equilibrium equations is demonstrated.
Elementary concepts of equilibrium and compatibility are then deployed to demonstrate
how to assemble the total stiffness matrix of a structure made up of an assembly of spring
elements by directly superimposing the stiffness matrices of the individual elements. The
term direct stiffness method refers to these techniques.
Simple illustration of imposing boundary conditions, and solving the resulting matrix
equations leading to nodal displacements and support reactions are presented. The prin-
ciple of minimum potential energy is introduced as a method to formulate element and
system equations relating forces and displacements. Some examples using this principles
are presented.

1.1 Linear spring element


A linear spring element is a mass less one dimensional mechanical model that resists
external forces by developing uniaxial stresses (either tension or compression). Such a
model is depicted in Figure 1.1. A spring is typically supported at one end, and an
axial load is applied at the other end. The applied load is directly proportional to the
displacement of the loaded end relative to the other end. If the other end is fixed, as is
the case here, relation between the force and the displacement of the free end is linear,
as shown in Figure 1.1c. The slope of the force-displacement line is called as the stiffness
of the spring, k. Stiffness can thus be defined as the force required to produce unit
displacement of the loaded end relative to the other end. Stiffness of a mechanical or a
structural element is a measure of its resistance to deform under the influence of external
actions.

4
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 5

B A P

(a)

Δ k
1
B A Δ
P
(b) (c)

Figure 1.1: Linear spring

The two ends of the spring, A and B, are called as its nodes. In order to completely
specify the deformed configuration of the spring, the displacements at the two ends of
the spring must be known. In other words at least two displacement components (in-
cluding the 0 displacement if one of the nodes is fixed) are needed to completely specify
the deformed configuration of the spring. Thus we say that the spring has two degrees of
freedom (DOF). Corresponding to each displacement component, there is a force compo-
nent. The independent coordinates along which the displacements (or forces) are specified
are called as DOF. The total number of DOF of a structural or a mechanical system is
the minimum number of independent coordinates required to completely specify the de-
formed configuration of the system. In this sense, DOF should not be associated with
displacements alone, but should be looked upon as coordinates along which both forces
and displacements are specified.

1.1.1 Element stiffness matrix


Let us now consider the spring isolated from the mechanical system (spring and the
support) as shown in Figure 1.2, and referred to as a spring element. The two nodes of
the spring are numbered as 1, and 2. The red arrows represent the 2 DOF of the spring,
at each of which both displacement and force are specified. Thus there are two force
components and two displacement components associated with the spring. We intend to
obtain a relationship between the forces and displacements in terms of the stiffness of the
spring, k.

k
u1, f1 u2, f2
1 2

Figure 1.2: Linear spring element

Invoking equilibrium of forces we obtain

f1 + f2 = 0
⇒ f1 = −f2 (1.1)
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 6

The constitutive behavior (force-deformation relation, or stress-stain relation) of the


spring is included in the definition of its stiffness, i.e.,

f2 = k(u2 − u1 )
⇒ f2 = −ku1 + ku2 (1.2)

Combining Equations 1.1 and 1.2 we obtain


f1 = ku1 − ku2 (1.3)
Equations 1.3 and 1.2 can be written in a matrix form as

    
f1 k −k u1
=
f2 −k k u2
⇒ {f } = [k] {u} (1.4)

This equation describes the relation between the forces and displacement of the spring
element. The vector of nodal forces {f } is called as element nodal force vector, and {u} is
called as the element nodal displacement vector. The square symmetric matrix [k] which
relates the forces and displacements is called as the element stiffness matrix. Owing to
Maxwell’s reciprocal law, the stiffness matrix of a structural or mechanical element is
always symmetric.

1.1.2 Boundary condition


The results obtained so far indicate that the displacements can be obtained as [k]−1 {f }.
However a meaningful absolute displacements of the nodes can be not obtained unless
some boundary conditions are specified. The boundary condition in this case can be the
displacement of one of the nodes. Mathematically, this is reflected by the singular nature
of the element stiffness matrix: the matrix [k] has a 0 determinant, and its inverse is
not defined. This means that the two equations defined by the matrix Equation 1.4 are
linearly dependent. In fact, if we consider the fact that f1 = −f2 , the two equations
are exactly the same, and therefore two unknown displacements can not be determined
independently. It is therefore essential to specify some boundary conditions on the matrix
equation. Consider the example when the node 1 is fixed leading to u1 = 0. This is called
as displacement boundary condition. Let the force at node 2 be known and equal to
P . This is called as force boundary condition. We can now apply these conditions to
Equation 1.4
    
f1 k −k 0
=
P −k k u2
which gives us u2 = P/k , as expected. This is equivalent to disregarding the equations
corresponding to those DOF where the displacements are specified as 0. In this case, DOF
1 is specified with 0 displacements, and thus removing row and column 1 of the stiffness
matrix, rows 1 of the force and displacement vector gives us a reduced set of equations (1
in this case) that can be solved. The force f1 can then be obtained from the first of the
two equations, i.e., f1 = −ku2 = −P . This satisfies the equilibrium of forces as expected.
We can thus conclude that if displacement boundary conditions are specified as 0, we can
obtain the reduced set of equations by removing those rows and columns from the stiffness
matrix, and those rows from the displacement and force vector which correspond to the
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 7

DOFs with specified 0 displacements. If on the other hand, the specified displacements
are non-zero, the above conclusion is not valid. For example, if node 1 is not fixed but
has some displacement δ, then the system of equation becomes
    
f1 k −k δ
=
P −k k u2
which gives us u2 = (P + kδ)/k , or P = k(u2 − δ) as expected. The force at the other
node is again given by f1 = −P , which must be the case to satisfy equilibrium of forces.
Note that at DOF 1, displacement is specified and force is the computed, while at DOF
2, force is specified and displacement is computed. It is incorrect to specify both force
and displacement at the same DOF. In this example, it would be incorrect to specify any
value of f1 other than −P as that would clearly violate equilibrium of forces.

1.2 An assembly of linear springs


Let us consider now an assembly of 2 springs as shown in Figure 1.3. The nodes are
indicated by red dots and numbered from 1 to 3. Forces P and Q are applied at nodes
2 and three respectively and node 1 is fixed. We intend to determine the unknown
displacements at nodes 2 and 3.

k1 k2
P Q

1 2 3

Figure 1.3: Linear spring element

1.2.1 Element stiffness matrices


Let us isolate each element separately and write the relation between the forces and
displacements at their nodes. The isolated springs are shown in Figure 1.4 along with the
nodal forces and displacements. Note that the force at DOF 2 carried by the first spring
is denoted as f21 and the force at the same DOF carried by the second spring is denoted
as f22 . When more than one elements share a common DOF, the displacements of all the
members at that DOF must be the same to imply continuity or compatibility. However,
the forces carried by all the members at the common DOF are generally not the same
and depend on the stiffness of the members, as we shall see shortly.
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 8

k1
f1, u1
f21, u2
1 2

k2
f22, u2
f3, u3
1 2

Figure 1.4: Linear spring element

With reference to Figure 1.4 and Equation 1.4 we can write the force displacement relation
for the first spring as     
f1 k1 −k1 u1
= (1.5)
f21 −k1 k1 u2
and that for the second spring
 as    
f22 k2 −k2 u2
= (1.6)
f3 −k2 k2 u3

1.2.2 Assembly of system equations


It is required to find a force displacement relation for the whole system (i.e., all the DOFs
of the spring assembly) so that the nodal displacements can be computed. For this we need
to assemble the force displacement relations (Equations 1.5and 1.6) of all the elements in
a proper manner. Such a procedure is called as assembly of system equations. Here we
will follow a simple and direct method to assemble the system equation from the element
equations. First we expand the equation for each element to include all the DOFs. In this
example, Equation 1.5 is expanded as
    
 f1  k1 −k1 0  u1 
f21 =  −k1 k1 0  u2 (1.7)
   
0 0 0 0 u3
and Equation 1.6 is expanded
 as   
 0  0 0 0  u1 
f22 =  0 k2 −k2  u2 (1.8)
   
f3 0 −k2 k2 u3
Now we can add
 theexpanded
 element Equations
  to obtain   
 F1  k1 −k1 0 0 0 0  u1 
F2 =  −k1 k1 0  +  0 k2 −k2  u2 (1.9)
   
F3 0 0 0 0 −k2 k2 u3
where F2 = f21 + f22 is the total force at DOF 2. This equation relates the displacements
of all the DOFs with the corresponding forces, and thus is called the system equation.
The sum of the square matrices inside the parenthesis is called as the system stiffness
matrix. The equation can be written as
{F } = [K] {U }
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 9

where uppercase letters are used to denote that we refer to the system equation rather
than the element equations. Thus the
 system stiffness matrix
 becomes
k1 −k1 0
[K] =  −k1 k1+ k2 −k2 
0 −k2 k2
which is again a square symmetric matrix. This matrix is also singular and appropriate
boundary conditions must me applied before the unknown displacements can be solved.

1.2.3 Boundary conditions


The displacement boundary condition is that u1 = 0.The force boundary conditions are
that F2 = P, and F3 = Q. Since the specified displacement at DOF 1 is 0, we can
remove the rows and columns 1 of the matrices and vectors in Equation 1.9 to apply the
appropriate boundary conditions. This results in
    
F2 k1 + k2 −k2 U2
= (1.10)
F3 −k2 k2 U3

1.2.4 Solution of unknown displacements


The unknown displacements can be solved from Equation 1.10 as

   −1      ( )
1 1 P +Q
U2 k1 + k2 −k2 P k1 k1 P k1
= = 1 1 1 = P +Q
U3 −k2 k2 Q k1 k1
+ k2
Q k1
+ kQ2

1.2.5 Solution of support reactions


The unknown reaction forces at the supports can be obtained from the unreduced system
equation (Equation 1.9) as

F1 = k1 U1 − k1 U2 + 0 × U3
 
P +Q
= −k1 = − (P + Q)
k1
which clearly satisfies the equilibrium of forces. This example shows the basic idea behind
the solution of systems which are an assembly of well-defined elements. Such systems are
also called as discrete systems. The general procedure adopted so far can be summarized
as
• Divide the system into several elements
• Formulate the force-displacement equations of each element separately, i.e., con-
struct element stiffness matrices
• Assemble the system equations from the element equations, i.e., assemble system
stiffness matrix
• Apply appropriate boundary conditions
• Solve the unknown displacements
• Solve the unknown reaction forces by using the displacements computed above
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 10

1.3 Stepped bar: a discretized continuous system


The problems we have considered so far involved discrete spring elements, defined with a
finite stiffness. In many practical problems, however, the discrete nature of the problem
is not as obvious, and the analyst needs to divide the problem of interest into discrete
elements, each of which are easily formulated. This is one of the basic principles behind
the finite element method. Let us consider a stepped bar as shown in Figure 1.5a which
is made of a material with Young’s modulus of elasticity E. The bar is fixed at the left
end while an axial load is applied at the free end. The lengths and area of cross section
of the two segments are as shown in Figure 1.5a.

1.3.1 Discretization
The difference in area of cross section of the two segments implies different stresses, and
therefore different strains in the two sections. As we anticipate the stress, and therefore
the strain to remain constant in each segment, it is obvious that the axial displacement at
each section varies linearly along the length. We therefore conclude that the displacement
throughout the bar can be determined if the displacement at the ends of each section we
known. We select nodes at the ends of each segment, and determine the nodal displace-
ments. The segments (elements) and nodes are shown in Figure 1.5b.

A1, E, L A2, E, L P 1 1 2 2 2 3

(a) (b)
y

i j
ui , f i A, E, L uj , fj
x
(c) (d)

Figure 1.5: A stepped bar

1.3.2 Formulation of element equations


A typical bar element is shown in Figure 1.5c. This generic element has nodes i, and j
with two degrees of freedom as shown in the figure. The displacement and force at node i
are denoted as ui and fi respectively. We need to establish a relation between the nodal
forces and the displacements. For the element, we know that the axial stress is given as
σx = fj /A. Since this is an uniaxial problem, the strain is given by ex = σx /E. From
kinematics we have
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 11

∂u
= ex
∂x
ˆuj ˆL
⇒ ∂u = ex ∂x
ui 0
⇒ uj − ui = ex L
fj L
⇒ = uj − ui
AE    
AE AE
⇒ fj = uj − ui (1.11)
L L

From equilibrium of forces we have    


AE AE
fi = −fj = − uj + ui (1.12)
L L
Combining Equations 1.11
 and1.12we get
  
fj AE 1 −1 uj
= (1.13)
fi L −1 1 ui
This equation resembles the element equation of a spring in Equation 1.4 with k = AE/L.
Thus a bar element carrying axial load only is quite similar to a linear spring.

1.3.3 Assembly of system equations


The assembly procedure follows exactly the same steps as in Section 1.2.2 with the spring
stiffnesses replaced by k1 = A1 E/L and k2 = A2 E/L.

1.3.4 Boundary condition


The displacement boundary condition is U1 = 0 and the force boundary condition is
F3 = P and F2 = 0. The system equation obtained as described above is
   AE −A1 E
 
 F1  1
L L
0  U1 
F2 =  −AL1 E AL1 E + AL2 E −AL2 E  U2 (1.14)
  −A2 E A2 E  
F3 0 L L
U3
Since DOF 1 is fixed, we remove the first equation from the system of Equations 1.14,
resulting in
    
0 E A1 + A2 −A2 U2
= (1.15)
P L −A2 A2 U3

1.3.5 Solution of unknown displacements


The unknown displacements can not be obtained by inverting Equation 1.15.

   −1    PL 
U2 L A1 + A2 −A2 0 A1 E
= = PL PL
U3 E −A2 A2 P A1 E
+ A2 E
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 12

1.3.6 Solution of unknown support reactions


The reaction force at the support is obtained from the unreduced system Equation 1.14
as

   
A1 E A1 L
F1 = U1 − U2 + (0) U3 = −P
L E

which satisfies the equilibrium of forces as expected.

1.3.7 Finding the displacement field


The nodal displacement can be interpolated within each element to find the displacement
field. In this example, a linearly varying displacement field is obtained for each element.
For element 1, assuming that the origin of coordinates coincides with the support we get
P
U (x) = x 0≤x≤L
A1 E
And for element 2, we use again a linear interpolation
PL P
U (x) = + (x − L) L≤x≤L
A1 E A2 E

1.3.8 Finding the strain field


The strain field can be obtained from the displacement field obtained above, i.e., ex (x) =
∂U (x)
∂x

P
ex (x) = 0≤x≤L
A1 E
P
ex (x) = L ≤ x ≤ 2L
A2 E
The results indicate that the strain is constant in each element, which is always the case
when the displacement varies linearly.

1.3.9 Finding the stress field


The stress field is determined from the strain field by using the constitutive relations, or
Hooke’s law. For the uniaxial case, we have the relation, σx (x) = Eex (x). Thus we obtain
constant stress field in each element: for element 1 σx (x) = P/A1 0 ≤ x ≤ L and for
element 2 σx (x) = P/A2 L ≤ x ≤ 2L.

1.4 Tapering bar: a continuous example


The previous example of stepped bar, although not so obviously discrete as the spring
assembly example, was easily discretized into two elements. The breakdown of the system
into elements was obvious due to the sudden change in area of cross section of the bar. In
many practical situations, such apparent discreteness is not inherent. In those situations,
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 13

the breakdown of a system into elements is often arbitrary. We consider such an example
in this section in the form of a tapered bar as shown in Figure 1.6a. The cross-sectional
area of the bar varies linearly from A1 at the left end to A2 at the right end.

1 2 3 i i+1 n n+1
A1 E, L A2 P 1 2 i n P

(a) (b)
y
A, E, l
i i+1
ui , f i i uj , fj
x
(c) (d)

Figure 1.6: A tapering bar

1.4.1 Discretization
We divide the tapering bar into n bars, each with a constant area as shown in Figure
1.6b. For simplicity let the length of each division be equal to l = L/n. It is obvious
that as n is increased, the subdivision of the bar becomes more and more accurate. The
selection of the number of elements n is arbitrary at this stage, and will be discussed
in more detail subsequently. Each element is now treated as a two node bar element as
discussed previously. The nodes are indicated by red circles, each numbered from 1 to
n + 1. The elements numbers are indicated inside circles enclosed inside the members.

1.4.2 Formulation of element equations, assembly of system equa-


tions, and solution
A typical element is shown in Figure 1.6c. The cross-sectional area varies linearly along
the member, and at any node i the area is given by
A1 − A2
Ai = A1 −
(i − 1)
n
The area of the member i is selected as the average area at node i and i + 1.
Ai + Ai+1
A =
2
2A1 − A1 −A
n
2
(i − 1 + i)
=
2
A1 − A2
= A1 − (2i − 1)
2n
Using the area of cross-section, the stiffness matrix of each element is formulated as in
Section 1.3.2. The assembly of system equation and the rest of the solution proceeds as
in Section 1.3.2. In this case, a computer program is used to perform the calculations.
The results are shown in Figure 1.7. The data used are A1 = 10; A2 = 1; P = 1; E = 1;
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 14

L = 1. The exact solution to this problem is plotted with the dashed black line given by
the following equation.    
−P L x A2 x
u= ln 1 − −
A1 − A2 L A1 L
As the number of elements is increased, the solution gets refined. It is noted that the
difference in the solution obtained by using 3 and 5 elements is not significant. This indi-
cates that increasing the number of element further is not expected to result in significant
improvement in the solution. The solution obtained by using 5 elements is fairly close
to the exact solution. If the analyst is not sure about how many elements to use, it is
useful to iteratively increase the number of elements until the solution is fairly stable and
insensitive to an increase in the number of elements.

0.35
n=1
n=2
0.3 n=3
n=5
Normalized displacment u

exact
0.25

0.2

0.15

0.1

0.05

0
0 0.2 0.4 0.6 0.8 1
x
Distance from support, L

Figure 1.7: Displacement along the length of the bar

The resulting strain-field is shown in Figure 1.8. The blue line represents the strain
obtained by using 5 elements. Note that strain within each element is constant as a
consequence of assuming a linear displacement field within an element. The exact solu-
tion is represented by the red curve. It is interesting to note that although the exact
displacement and that obtained using 5 elements were fairly similar, the difference in
corresponding strain field is more prominent. This is usually the case with finite element
results. Generally a more detailed approximation of displacement field is needed to es-
tablish a reliable strain field. If the displacement field in an element is linear, as is the
case in this example, more elements are needed to obtain better accuracy in the strain
field, and consequently the stress field.
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 15

1
n=5
0.9 exaxt

0.8

0.7

Strain ex
0.6

0.5

0.4

0.3

0.2

0.1
0 0.2 0.4 0.6 0.8 1
x
Distance from support, L

Figure 1.8: Strain along the length of the bar

1.5 Potential energy approach


The formulation of element equations in the preceding sections was based on equilibrium
equations and constitutive laws. An alternative method often used to derive element
equations and the stiffness matrix for an element is based on the principle of minimum
potential energy. This method has the advantage of being more general and adaptable
to the determination of element equations for elements with large number of degrees
of freedom. In subsequent chapters, this principle will be used to establish the basic
formulation of the finite element method for an elastic continuum. In this section, only
simple examples are presented in order to introduce the basic concepts.

1.5.1 Principle of stationary potential energy


The principle of stationary potential energy is a special case of the virtual work principle
and can be stated as

Of all admissible displacements of a continuous elastic body, those that satisfy


the equations of equilibrium make the potential energy stationary with respect
to small admissible variations of displacements. If the stationary condition is
a relative minimum, the equilibrium state is stable.

To explain this principle, we need first to understand the concepts of potential energy and
stationary value of a function. Total potential energy is defined as the sum of the internal
strain energy U and the potential energy of the external forces W . Strain energy is the
capacity of internal stresses to do work through strains. Potential energy of external forces
is the capacity of forces such as body forces, surface forces, and applied nodal forces to
do work through deformation of the structure. These concepts are illustrated for a simple
discrete system in this section, and a more detailed treatment for elastic continuum is
deferred until later chapters.
Consider the linear spring element of Figure 1.1a. We know that the spring force F
is related to the displacement of the free end u by the equation F = ku. If the spring
is deformed by a small amount du, the differential work done by the internal forces gives
the differential strain energy
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 16

dU = kudu (1.16)
The total strain energy is deforming the spring from u = 0 to u = ∆ is obtained as
ˆ∆ ˆ∆
k∆2
U = dU = k udu =
2
0 0

The work done by the external force P is P ∆. This work done is taken as negative
potential because the potential energy of the external force is lost when the work is
done,.i.e,
W = −P ∆
The potential energy is now given by
k∆2
Πp = U + W = − P∆ (1.17)
2
The principles of variational calculus can be used to minimize the potential energy. In this
section, however, we use simple differential calculus, i.e., if ∆e represents the equilibrium
displacement of the spring then
∂Πp
= 0
∂∆
P
⇒ ∆e =
k
The variation of potential energy with admissible displacement is shown in Figure 1.9.
Note that the potential energy is minimum corresponding to the displacement that satisfies
equilibrium equation.

U
W
Πp
Potential

∆e

Displcament, ∆

Figure 1.9: Graphical interpretation of the minimum potential energy principle

1.5.2 Stiffness matrix of a spring


In the previous example, the spring was supported at one end, and a single DOF described
the admissible displacement of the spring. In such cases, the minimization of potential
energy with respect to the single displacement results in one equation that gives the
equilibrium displacement of the spring. If there are more than one DOF, then the potential
energy should be minimized with respect to the displacements at all the DOF, resulting
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 17

in as many equations as there are DOF in the element. This results in a s system of
equations known as the element equation containing the element stiffness matrix. To
illustrate the discussion let us consider now the spring element of Figure 1.2. There are
two DOF in this case. The potential energy is computed in the following steps

1
U = k (u2 − u1 )2
2
W = −f1 u1 − f2 u2
1
Πp = U + W = k (u2 − u1 )2 − f1 u1 − f2 u2 (1.18)
2
For stable equilibrium the potential energy must be minimized with respect to both u1
and u2 . Considering minimization with respect to u1 we obtain
∂Πp
= 0
∂u1
⇒ ku1 − ku2 − f1 = 0
⇒ f1 = ku1 − ku2 (1.19)

Considering minimization with respect to u2 we obtain


∂Πp
= 0
∂u2
⇒ f2 = −ku1 + ku2 (1.20)

Combining Equations 1.19 and


 1.20 we obtain  
f1 k −k u1
=
f2 −k k u2
which gives the element equation similar to the one in Equation 1.4.

1.5.3 Assembly of linear springs


Let us now consider the assembly of linear springs shown in Figure 1.3. The element
equations for the two springs can be formulated by using the equilibrium method or the
minimum potential energy method as discussed previously. Then the element equations
can be assembled directly as described in Section 1.2.2. It is also possible to apply the
principle of minimum potential energy to the whole assembly thus obtaining the system
equations at once. Such an approach discussed here to illustrate that the assembly of
element equations is in fact equivalent to minimization of the total potential energy of
the system with respect to all the displacement DOF.
The potential energy of the spring assembly is computed in the following steps
1 1
U = k1 (U2 − U1 )2 + k2 (U3 − U2 )2
2 2
W = −F1 U1 − P U2 − QU3

where F1 is the support reaction force.


1 1
Πp = k1 (U2 − U1 )2 + k2 (U3 − U2 )2 − F1 U1 − F2 U2 − F3 U3
2 2
Minimizing the potential energy with respect to U1 we get
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 18

∂Πp
= 0
∂u1
⇒ k1 U1 − k1 U2 − F1 = 0
F1 = k1 U1 − k1 U2 + 0 (U3 ) (1.21)

Similarly proceeding with U2 we get


∂Πp
= 0
∂U2
⇒ k1 (U2 − U1 ) − k2 (U3 − U2 ) = F2
⇒ F2 = −k1 U1 + (k1 + k2 ) U2 − k2 U3 (1.22)

Similarly proceeding with U3 we get


∂Πp
= 0
∂U3
⇒ k2 (U3 − U2 ) − F3 = 0
⇒ F3 = (0) U1 − k2 U2 + k2 U3 (1.23)

Combining Equations 
1.21, 1.22,
 and 1.23 we obtain  
 F1  k1 −k1 0  U1 
F2 =  −k1 k1 + k2 −k2  U2
   
F3 0 −k2 k2 U3
which is the same as Equation 1.9. Now the boundary conditions U1 = 0, F2 = P , and
F3 = Q can be applied and the reduced system of equations can be solved to find the
unknown displacements. This example shows that the principle of minimum potential
energy can be applied in two ways

• applying the principle separately to each element to obtain the element equations
which are assembled to obtain system equations

• applying the principle to the whole system to obtain the system equations

In either case, the same set of system equations are obtained. The first approach is easier
to implement and is preferred in finite element calculations.

1.5.4 Example of a truss structure


Consider the truss structure shown in Figure 1.10. A truss structure consists of prismatic
bars which are connected by pinned joints. Pinned joints can not transfer any moment.
A truss member can carry only axial forces resulting from concentrated loads applied at
the joints. No load or moment is applied on the member itself. The truss in this example
consists of 3 members as marked. The nodes are marked with red circles, 4 altogether.
The forces and displacements are to be specified in the XY coordinate plane the origin
of which passes through node 1. A concentrated load of 1 kN is applied at node 1 as
indicated in the figure. All the members have same area A and Young’s modulus E, and
AE = 2 × 105 N. The supports at nodes 2, 3, and 4 are pinned meaning that they cannot
translate (move in X or Y direction). It is required to find the displacement at node 1,
and support reactions. The problem can be solved by using the principle of minimum
potential energy. In the following two approaches of solution are discussed.
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 19
x

2 3

1 2 1m

1 4
3 X
1m

1kN

Figure 1.10: A truss structure

1.5.4.1 Solution by applying minimum potential energy principle to the whole


structure
In this approach, the principle of minimum potential energy is applied to the whole
structure to obtain the system equations. The DOF are selected starting at node 1,
where X-displacement is selected as DOF 1 and Y-displacement as DOF 2. The scheme
proceeds to the other nodes. For any node i the displacements in the X and Y direction
are D2i−1 and D2i respectively. Since a truss element can only deform parallel to its axis,
if the coordinate is parallel to the axis of the truss element, only two DOF, one at each
node, is sufficient. On the other hand, if a truss element is inclined, the axial displacement
at each node has two components in the XY coordinate system, and therefore there are
2 DOF at each node, and 4 DOF in total. The XY coordinate system in this example is
the global coordinate system, which is used as a common coordinate to specify forces and
displacements in all the members.
For any member which is arbitrary oriented in the XY plane the potential energy is
expressed as a function of its nodal displacements in the following. Consider a generic
member with nodes i and j as shown in Figure 1.11. The member is inclined at an
angle α from the global X axis. A coordinate system xy is selected in such a way that
x is parallel to the member, and has its origin at node i. Such a coordinate system
which is specific to a member is called as the local or element coordinate system. Local
coordinate system makes the formulation of element equations easier. In this text we use
lowercase and uppercase letters for local and global coordinate system respectively. The
coordinates of the nodes with respect to the global coordinate system are (Xi , Yi ) and
(Xj , Yj ). In the element coordinate system, the truss member has only 2 DOF, namely
the displacement parallel to x axis at nodes i and j, which are denoted as di and dj . In
the global coordinate system, there are 4 DOF, namely D2i , D2i−1 , D2j , and D2j−1 . We
intend to find an expression for the strain energy of the bar in terms of the forces and
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 20

displacements in the global coordinate system.

x
D2j

dj
D2j-1
j(Xj,Yj)
D2i

y
α
Y D2i-1
di i(Xi,Yi)

Figure 1.11: A generic truss element

The member is analogousq to a spring with stiffness k = AE/L with the length of the
member given byL = (X2 − X1 )2 + (Y2 − Y1 )2 . Therefore its strain energy is given by
 
1 AE
U= (dj − di )2 (1.24)
2 L
Thus strain energy is easily formulated in local coordinate system. However, we wish to
cast the final equations in terms of the global coordinate system. This is achieved by the
using the simple geometrical relations
di = D2i−1 cos α + D2i sin α (1.25)
and a similar equation for node j. Note that sin α = (Y2 −Y1 )/L, and cos α = (X2 −X1 )/L.
Also note that D3 = D4 = D5 = D6 = D7 = D8 = 0. Using these equations we compute
the strain energies for all the elements separately as follows

Element 1
 
1 AE
U1 = (d2 − d1 )2
2 1
 
AE
= [(D3 cos 90° + D4 sin 90°) − (D1 cos 90° + D2 sin 90°)]2
2
 
AE 
= D22
2

Element 2
 
1 AE
U2 = √ (d3 − d1 )2
2 2
 
AE
= √ [(D5 cos 45° + D6 sin 45°) − (D1 cos 45° + D2 sin 45°)]2
2 2
 
AE 
= √ D12 + 2D1 D2 + D22
4 2
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 21

Element 3
 
1 AE 2
U3 = d24 − d21
2 1
 
AE
= [(D7 cos 0° + D8 sin 0°) − (D1 cos 0° + D2 sin 0°)]2
2
 
AE 
= D12
2

And the total strain energy is given by


U = U1 + U2 + U3
The potential of external work by a force of 1kN is given by
W = −(−1000)D2 = 1000D2
And the potential energy is therefore
Πp = U1 + U2 + U3 + 1000D2
which is a function of the nodal displacements D1 and D2 . Minimizing the potential energy
with respect to each of the displacements gives us 2 equations as follows
∂Πp
= 0
∂D
    1
AE AE
⇒ √ (2D1 + 2D2 ) + 2D1 = 0
4 2 2
   
AE
⇒ D1 √ + AE +D2 2AE

2
= 0 (1.26)
2 2

∂Πp
= 0
∂D2
AE AE
⇒ AED2 + √ D1 + √ D2 + 1000 = 0
2 2 2 2
   
AE AE
⇒ D1 √ + D2 AE + √ = −1000 (1.27)
2 2 2 2
Combining Equations 1.26 and 1.27 we get
     
D1 1.354 0.354 0
AE = (1.28)
D2 0.354 1.354 −1000
    −1  
D1 1 1.354 0.354 0
⇒ =
D2 AE 0.354 1.354 −1000
 
0.001
= m (1.29)
−0.004

Indicating that node 1 moves to the right by 1 mm and downwards by 4 mm.


CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 22

1.5.4.2 Solution by assembly of element stiffness matrices


In this method we formulate the stiffness matrix of each element separately and assemble
them to obtain system stiffness matrix. In the element coordinate system, the element
equation of a generic element as that shown in Figure 1.11 is the same as that derived in
Section 1.3.2, i.e.,     
fi AE 1 −1 di
=
fj L −1 1 dj
where the element stiffness matrix is given as 
0 AE 1 −1
k =
L −1 1
Since the members are oriented in different manner, the element coordinate system is
different for different member and we can not assemble the system equations by simply
adding the element equations like we did in Section 1.3.3. We could add the element
equations if they were expressed in the global coordinate system. We seek to find the
stiffness matrix of the element in the global coordinate system. The relation between the
nodal displacement in the local and global coordinate system is obtained from geometry
as

di = D2i−1 cos α + D2i sin α


= lD2i−1 + mD2i (1.30)

where l = cos α and m = sin α = cos(90°-α) are the direction cosines of the local x axis.
Similarly
dj = lD2j−1 + mD2j (1.31)
Combining Equations ?? and 1.31 in a matrix form we get
 
   

D2i−1 


di l m 0 0 D2i
=
dj 0 0 l m   D2j−1 

 
D2j
n 0o
d = [T ] {d} (1.32)

where {d} represents the element nodal displacements in global coordinate system. Note
that lowercase letter is used to represent element, where as {D}would representthe nodal
0
displacement of the the whole system in global coordinate system. The vector d is the
nodal displacement vector of the element in local coordinate system. Note that the prime
is used to indicate local coordinate system. The matrix [T ] is the transformation matrix
that converts element displacement vector from global coordinates to local coordinates.
We now seek the relationship between the element stiffness matrix in global and local
coordinates. To achieve this make use of the fact that the work done by forces in both
coordinate system must be the same, as energy is invariant to coordinate system. The
work done in local coordinate system is
n 0 oT n 0 o
w= f d
 0
where f is the element nodal force vector in local coordinates. Similarly the work done
in global coordinate system is
w = {f }T {d}
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 23

Equating the work done we obtain


n 0 oT n 0 o
f d = {f }T {d}
h 0 i n 0 oT n 0 o
⇒ k d d = ([k] {d})T {d}
h 0 i T
⇒ k [T ] {d} [T ] {d} = ([k] {d})T {d}
h 0 iT
{d}T [T ]T k [T ] {d} = {d}T [k]T {d}
h 0 iT
⇒ [T ]T k [T ] = [k]T
h 0i
T
⇒ [k] = [T ] k [T ] (1.33)

which gives us the transformation equation between element stiffness matrices in local and
global coordinate systems. Once the element matrices are found in global coordinates they
can be simply added according to the DOF of each element to find the system matrices.
This procedure is illustrated next
Initialize the system stiffness matrix with zeros
 
0 0 0 0 0 0 0 0
 0 0 0 0 0 0 0 0 
 
 0 0 0 0 0 0 0 0 
 
 0 0 0 0 0 0 0 0 
[K] = 
 0 0 0 0 0 0 0 0 
 (1.34)
 
 0 0 0 0 0 0 0 0 
 
 0 0 0 0 0 0 0 0 
0 0 0 0 0 0 0 0

Find the stiffness matrix of each element and add it to the system stiffness matrix based
on its DOFs

Element 1
h 0i  
AE 1 −1
k =
1 −1 1
 
0 1 0 0
[T ] =
0 0 0 1
 
0 0   
AE  
 1 0  1 −1 0 1 0 0
[k] = 
1 0 0  −1 1 0 0 0 1
0 1
 
  0 0 0 0
AE   0 1 0 −1 

=
1  0 0 0 0 
0 −1 0 1

Element 1 corresponds to DOFs 1, 2, 3, and 4, therefore its stiffness matrix is added to


rows and columns 1, 2, 3, 4 of the system matrix, which gets updated as
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 24

 
0 0 0 0 0 0 0 0
 0 1 0 −1 0 0 0 0 
 
 0 0 0 0 0 0 0 0 
 
 0 −1 0 1 0 0 0 0 
[K] = AE 

 (1.35)
 0 0 0 0 0 0 0 0 
 0 0 0 0 0 0 0 0 
 
 0 0 0 0 0 0 0 0 
0 0 0 0 0 0 0 0

Element 2
h 0i  
AE 1 −1
k = √
2 −1 1
" #
√1 √1 0 0
2 2
T =
0 0 √12 √12
h 0i
[k] = [T ]T k [T ]
 
0.354 0.354 −0.354 −0.354
 0.354 0.354 −0.354 −0.354 
= 
 −0.354 −0.354 0.354

0.354 
−0.354 −0.354 0.354 0.354
Element 2 corresponds to DOF 1, 2, 5 and 6, and hence its stiffness matrix is added to
rows and columns 1, 2, 5, and 6 of the system matrix of Equation 1.35, which becomes
updated as
 
0.354 0.354 0 0 −0.354 −0.354 0 0
 0.354 1.354 0 −1 −0.354 −0.354 0 0 
 
 0 0 0 0 0 0 0 0 
 
 0 1 0 −1 0 0 0 0 

[K] = AE  
−0.354 −0.354 0 0 0.354 0.354 0 0 
 
 −0.354 −0.354 0 0 0.354 0.354 0 0 
 
 0 0 0 0 0 0 0 0 
0 0 0 0 0 0 0 0

Element 3
h 0i  
AE 1 −1
k =
1 −1 1
 
1 0 0 0
[T ] =
0 0 1 0
 
1 0 −1 0
 0 0 0 0 
[k] = 
 −1 0

1 0 
0 0 0 0
Element 3 corresponds to DOF 1, 2, 7, and 8 and its stiffness matrix is added to these
rows and columns of the system matrix, which is updated as
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 25

 
1.354 0.354 0 0 −0.354 −1 0 0
 0.354 1.354 0 −1 −0.354 −0.354 0 0 
 
 0 0 0 0 0 0 0 0 
 
 0 −1 0 1 0 0 0 0 
[K] = AE 


 −0.354 −0.354 0 0 0.354 0.354 0 0 
 −0.354 −0.354 0 0 0.354 0.354 0 0 
 
 −1 0 0 0 0 1 0 0 
0 0 0 0 0 0 0 0
This gives the final stiffness matrix of the system, which is singular due to the fact that
boundary conditions have not be applied. The displacement boundary condition is that
D3 = D4 = D5 = D6 = D7 = D8 = 0. Thus we remove the rows and columns 3, 4, 5, 6,
7, and 8 from the system stiffness matrix,resulting in a reduced
 stiffness matrix
1.354 0.354
[Kr ] = AE
0.354 1.354
And the reduced system of equation
 is  
F1 D1
= [Kr ]
F2 D2
now we apply the force boundary
 condition, 0 and
 F1 =   obtaining
 F2 = −1000,
1.354 0.354 D1 0
AE =
0.354 1.354 D2 −1000
which is exactly the same as Equation 1.28. In order to find the support reactions we can
use the unreduced system equation, i.e.,
   

 D 
  0 
  
1

  
 


 D 
2  
 −1000 


 
 
 


 0 
 
 0 

   
0 793
{F } = = [K] {D} = [K] =

 0   
 207  

 
 
 


 0 
 
 207 


  
  


 0 
 
 −207 

   
0 0

The support reactions are shown in Figure 1.12. Note that the equilibrium of forces is
satisfied, for example

ΣFX = F1 + F3 + F5 + F7 = 0
ΣFY = F2 + F4 + F6 + F8 = −1000 + 793 + 207 + 0 = 0

It can also be verified that the sum of moments about any point in the structure is 0.
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 26

F4= 793 F8= 207

F7= 207
F3= 0

1 2

F8= 0

F7= -207
3
F1= 0

F2= -1000

Figure 1.12: Support reactions and applied forces

This example illustrates in detail the procedures involved in solving a structural mechanics
problem by using the Finite Element Method. A detailed definition of the method, and
formal derivations are postponed until subsequent chapters. It is highlighted that applying
solution procedure of Section 1.5.4.1 seems shorter and easier in this example than that of
Section 1.5.4.2. On the other hand, the latter procedure follows a systematic set of steps
which can be easily coded in a computer. In addition, the latter procedure is readily
available to compute support reactions, while the former needs additional calculations
once the unknown displacements are determined. It is therefore customary to use the
latter procedure in practical applications.

1.6 Summary
This chapter illustrated the process of dividing a structural system into separate elements,
and formulating their force deformation relations. This can be achieved either by using
equilibrium of forces at the nodes of the elements, or by applying the principle of minimum
potential energy. The principle of minimum potential energy is a powerful tool which can
be applied both at the element and the structure level. To formulate element equations
we expressed the potential energy of the element as a function of unknown displacements
Πp = f (d1 , d2 , ....., dN ) where N stands for the number of DOF of the element. Mini-
mizing the potential energy with respect to each unknown displacement results in a set
of N equations, relating the element displacements and forces. This set of equations is
called as element equations which contains the element nodal force vector, element nodal
displacement vector, and the element stiffness matrix. Stiffness matrices of the elements
can be combined together after proper coordinate transformation as necessary, to obtain
the stiffness matrix of the structure. Finally proper displacement and force boundary
conditions are applied, and unknown displacements are determined by solving matrix
equations.
Chapter 2

Finite element formulation of elastic


continuum

2.1 Introduction
The examples dealt with in Chapter 1were mainly discrete in nature, and the expression
for potential energy of the elements were readily available in terms of nodal displacements.
This is, however, not the case for an elastic continuum. For an elastic continuum, the
potential energy is a function of the displacement field, and elastic properties of the ma-
terial. In such cases we divide the continuum into small elements with two or more nodes.
By assuming a certain pattern of displacement variation within the element, it is possible
to express the displacement field in terms of the displacements of the element nodes. This
allows us to express the potential energy of the element in terms of the displacements
of a selected finite number of nodes in the element. The principle of minimum potential
energy can then be applied to obtain a relationship between the displacements and forces
at the nodes. In this chapter we review some results from elastic theory of a continuum,
and formally derive finite element form of the equilibrium equations.

2.2 Strain tensor


Consider an elastic continuum of arbitrary shape in a three dimensional space as shown in
Figure 2.1. The coordinate system being considered is a right-handed 3D Cartesian type
(Z axis coming out of the plane of the paper). The displacement field within the elastic


body is denoted by U (X, Y, Z) and its components parallel to the coordinate axes are
UX , UY , and UZ which are all functions of coordinates X , Y , and Z. The concentrate
forces (both forces and moments/torques) acting on the elastic body are represented by


P , a vector in the three dimensional space. The distributed traction (surface force) over


a certain surface ST is denoted by T , which is measured in the units of force per unit
area. The body forces distributed throughout the volume of the elastic body is denoted


by B and is measured in units of force per unit volume. Over a certain surface SU , the
elastic body is supported in a specified manner, i.e., the displacement field in this surface
is known.

27
CHAPTER 2. FINITE ELEMENT FORMULATION OF ELASTIC CONTINUUM 28


T

P
ST


B
Y
SU

Figure 2.1: Elastic continuum in a three dimensional space

Kinematics is concerned with a description of the relationship between the displacement


and strain fields within the elastic body. For small displacements, the strain tensor is
related to the displacement field by the following
 equation

1 ∂Ui ∂Uj
eij = + (2.1)
2 ∂Xj ∂Xi
with i , j representing the indices corresponding to the X, Y , and Z axes. The second
order strain tensor is symmetric, and only 6 of its 9 components are independent. The
independent components are given as
∂UX
eXX =
∂X
∂UY
eY Y =
∂Y
∂UZ
eZZ =
∂Z
∂UX ∂UY
γXY = 2eXY = +
∂Y ∂X
∂UY ∂UZ
γXY = 2eY Z = +
∂Z ∂Y
∂UZ ∂UX
γZX = 2eZX = +
∂X ∂Z
The strain components eXX , eY Y , and eZZ are called the normal strains parallel to the X
, Y , and Z directions. The strain components eXY = eY X , eY Z = eZY , and eZX = eXZ
are the shear strains in the XY , Y Z, and XZ planes respectively. The engineering
CHAPTER 2. FINITE ELEMENT FORMULATION OF ELASTIC CONTINUUM 29

shear strain γ = 2e represents the rotation due to shear deformation in a plane. The
strain-displacement relation can be written in a matrix form as
   ∂ 

 e XX 
 ∂X
0 0

 
  0

 e Y Y 
 

∂Y
0 
 
   ∂   Ux 
eZZ 0 0 ∂Z 
=  UY

 γXY   
∂ ∂
0   

 
 
∂Y ∂X
∂ ∂  UZ

 γ  0
 YZ   ∂
∂Z ∂Y

γZX ∂Z
0 ∂X
⇒ {e} = [L] {U } (2.2)

where {e} is called the strain vector. The matrix of differential operators, [L] is called as
the differential operator matrix, and {U } is the displacement vector. It should be noted
that both {e}and {U } are a function of position, (X, Y, Z) and thus vector field quanti-
ties, more suitably denoted as {e(X, Y, Z)} and {U (X, Y, Z)} . However for notational
simplicity we use the symbols {e}and {U }.

2.3 Stress tensor


The stress tensor represents the components of stress acting on the faces of a infinitesimal
cube with its sides parallel to the coordinate axes. The stress tensor, like the strain
tensor, is a second order tensor consisting of 9 components, as shown in Figure 2.2. In
order to define the state of stress at a point, consider an infinitesimal cube around the
point. The sides of the cube are parallel to the coordinate axes. The right face of the cube
is normal to the positive X axis and is termed as the positive X plane. Similarly, the top
and the front faces are termed the positive Y and positive Z planes, respectively. The
left, bottom, and back faces are the negative X, Y , and Z faces , respectively. Consider
the stress components acting on the positive X face. The three components of stress on
this face are denoted as σXX , σXY , and σXZ . The first index in the subscript is X to
denote that the stress components correspond to the positive X plane. The second index
corresponds to the direction of the stress. For example, σXY represents stress on the X
plane and parallel to the positive X direction. The stress component perpendicular to the
face is called as the normal stress. The stress components parallel to the face are called
as shear stress. For example, on the X face, σXX is the normal stress while σXY and
σXZ are shear stresses. The stress components in the positive Y and positive Z faces are
defined accordingly. The convention is that a stress acting on a positive face, and parallel
to one of the positive coordinate axes is considered as positive. With this convention, on
the left face, the positive σXX would be directed in the negative X direction.
CHAPTER 2. FINITE ELEMENT FORMULATION OF ELASTIC CONTINUUM 30

Y
 YY

 YX
 YZ  XY
 ZY
 ZX  XX
 ZZ  XZ
X
Z
Figure 2.2: Stress components on the faces of a cube with sides parallel to the coordinate
axes.

For equilibrium, the shear stresses satisfy the conditions σXY = σY X , σY Z = σZY and
σZX = σXZ . Therefore, of the nine stress components, only 6 are independent. These
independent stress components can be grouped in a vector called as stress vector. In the
following we use the symbol τ for shear stresses, for example, τXY = σXY , and so on. The
stress vector is therefore  

 σ XX 


 


 σ YY 
 
σZZ
{σ} = (2.3)

 τXY  

 


 τ 
 YZ  
τZX

2.4 Constitutive relation


For a linearly elastic, isotropic, homogeneous material, the relation between stress and
strain is given by the generalized Hooke’s law. In index notations, generalized Hooke’s
law can be written as

σij = λekk δij + 2µeij


where, σij is the stress tensor; eij is the strain tensor; ekk = eXX + eY Y + eZZ is the volu-
metric strain; λ and µ are elastic properties of the material known as Lamé’s parameters;
and δij is the Kronecker’s delta defined as δij = 1 for i = j and δij = 0 otherwise. The
equation in index notations can be expanded to obtain the following six equations
CHAPTER 2. FINITE ELEMENT FORMULATION OF ELASTIC CONTINUUM 31

σXX = λ (eXX + eY Y + eZZ ) + 2µeXX


σY Y = λ (eXX + eY Y + eZZ ) + 2µeY Y
σZZ = λ (eXX + eY Y + eZZ ) + 2µeZZ
τXY = 2µeXY = µγXY
τY Z = 2µeY Z = µγY Z
τZX = 2µeZX = µγZX

Rewriting the equations in matrix form gives us


    

 σ XX 
 λ + 2µ λ λ 0 0 0 
 eXX 


 
   


 σY Y 
  λ λ + 2µ λ 0 0 0  eY Y 

    
σZZ  λ λ λ + 2µ 0 0 0  eZZ
=  

 τ 
XY   0 0 0 µ 0 0  γXY 


 
   


 τ  0 0 0 0 µ 0  γY Z 
 YZ   
 

τZX 0 0 0 0 0 µ γZX
⇒ {σ} = [E] {e} (2.4)

where the matrix [E] which relates the strain vector to the stress vector is known as the
elasticity matrix. Elasticity matrix is a function of the elastic properties of the material,
λ and µ which are related to the more familiar engineering properties, namely Young’s
modulus E , and Poisson’s ratio ν by the following equations.
E
µ =
2 (1 + ν)

λ =
(1 + ν) (1 − 2ν)

Note that µ is also commonly known as the shear modulus of elasticity, and many authors
denote it with the symbol G.

2.5 Elastic strain energy


Elastic strain energy represents the work done by the internal stresses in undergoing elastic
deformation. Elastic strain energy density (energy per unit volume) at a point is related
to the stress and the strain tensor at that point. Using the summation convention, the
strain energy density is given by
1
π = σij eij (2.5)
2
On expanding the summation indices we obtain
CHAPTER 2. FINITE ELEMENT FORMULATION OF ELASTIC CONTINUUM 32

1
π = (σXX eXX + σY Y eY Y + σZZ eZZ + 2τXY eXY + 2τY Z eY Z + 2τZX eZX )
2
1
= (σXX eXX + σY Y eY Y + σZZ eZZ + τXY γXY + τY Z γY Z + τZX γZX )
2  
 σXX 
 

 


 σ Y Y 

1   
σZZ
= eXX eY Y eZZ γXY γY Z γZX
2 
 τXY  

 


 τ Y Z 

 
τZX
1
= {e}T {σ} (2.6)
2
Combining this with Equations 2.2 and 2.4 we get
1
π = {e}T [E] {e}
2
1
= ([L] {U })T [E] [L] {U }
2
1
= {U }T [L]T [E] [L] {U }
2
The total strain energy is obtained by integrating π over the entire volume of the elastic
body. ˆ ˆ
1
Π = πdV = {U }T [L]T [E] [L] {U } dV (2.7)
2
V V

2.6 Work done by external forces


The work done by the surface traction components Tx , Ty , and Tz acting on a small surface
area dS in undergoing displacements UX , UY , and UZ is obtained as

dW T = −UX TX dS − U TY dS− Uz TZ dS


   TX 
= − UX UY UZ TY dS
 
TZ
= − {U }T {T } dS

and the total work done by the surface traction is obtained by integrating over the surface
ST ˆ
W = − {U }T {T } dS
T
(2.8)
ST

The work done by body force components BX , BY , and BZ acting on a small volume dV
in undergoing displacements UX , UY , and UZ is given by
 
   BX 
dW B = − UX UY UZ BY dV
 
BZ
= − {U }T {B} dV
CHAPTER 2. FINITE ELEMENT FORMULATION OF ELASTIC CONTINUUM 33

and the total work done by the body forces is obtained by integrating over the volume of
the elastic body. ˆ
W = − {U }T {B} dV
B

The work done by all the external forces is then

W = WˆT + W B ˆ
T
= − {U } {B} dV − {U }T {T } dS (2.9)
V ST

2.7 Potential energy


The potential energy of the elastic body is obtained as

Πp = Π + W
ˆ ˆ ˆ
1
= {U } [L] [E] [L] {U } dV − {U } {B} dV − {U }T {T } dS (2.10)
T T T
2
V V ST

Through this expression, the potential energy is given as a function of the displacement
field.

2.8 Formulation of element equations


The potential energy expression of Equation 2.10 involves a vector displacement field
which is a continuous function of the spatial coordinates. A complete description of the
displacement field thus requires the knowledge of displacement at each and every point
inside the continuum. Thus a solution is difficult, and often impossible to obtain. We
seek for an approximate solution to the problem in the form of a discrete displacement
field, which means the displacements being specified at a finite number of points in the
region of interest. One natural way of achieving this is to divide the elastic body into a
finite number of smaller bodies. Such smaller bodies are called as elements of the elastic
body. Each element is assigned two or more nodes, and the displacements at the nodes are
treated as unknown quantities. The nodal displacements of each element, when combined
together, gives an approximation to the displacement field of the entire body.
Let us now consider a small element inside an elastic body. The element consists of a
finite number of nodes. The unknown displacement at the nodes are collected in a vector
{d}. Let the displacement field within the element be denoted by {u(X, Y, Z)} . We seek
to express the displacement field in terms of the nodal displacement vector {d}. To achieve
this, we assume a certain pattern of displacement variation within the element, such as
linear variation, quadratic variation, etc. If a proper shape of displacement is assumed, the
displacement at the nodes can be interpolated within the element to obtain the complete
displacement field for the element. Mathematically, the interpolation operation can be
written as
{u(X, Y, Z)} = [N (X, Y, Z)] {d} (2.11)
where the matrix [N (X, Y, Z)] is a continuous function of the spatial coordinates, and
depends on the assumed displacement variation within the element. Such a matrix is
CHAPTER 2. FINITE ELEMENT FORMULATION OF ELASTIC CONTINUUM 34

known as shape function matrix. A more detailed treatment of the shape function matrix
is presented in the subsequent chapters, and it suffices for the developments here, to
think of it as an interpolation operator. The interpolation operation provides us with a
continuous field in terms of the values of the field at a discrete number of points. For
notational simplicity we simply write {u} = [N ] {d} with the understanding that both
{u} and [N ] are continuous functions of the spatial coordinates. The size of the element
nodal displacement vector {d} is the number of degrees of freedom of the element, denoted
here as n .
The elastic strain energy of the element can be computed from Equation 2.7 as

ˆ
1
Π = {u}T [L]T [E] [L] {u} dV
2
ˆ
V
1
= {d}T [N ]T [L]T [E] [L] [N ] {d} dV
2
V
 
ˆ
1
= {d}T  [N ]T [L]T [E] [L] [N ] dV  {d}
2
V

We define as strain matrix [S] = [L][N ]. It is called as strain matrix because it gives the
strain field in terms of the nodal displacement vector, i.e., {e} = [L] {u} = [L] [N ] {d} =
[S] {d}. The strain energy thus becomes 
ˆ
1
Π = {d}T  [S]T [E] [S] dV  {d}
2
V

The integral inside the brackets results in a matrix which is completely defined by the
elasticity matrix, and the shape function matrix, and is called as the stiffness matrix of
the element, [k]. Thus
1
Π = {d}T [k] {d} (2.12)
2
The work done by the external body and surface forces is obtained from Equation 2.9 as
ˆ ˆ
W = − {u} {B} dV − {u}T {T } dS
T

Ve Se
ˆ ˆ
T T
=− {d} [N ] {B} dV − {d}T [N ]T {T } dS
Ve e S
 
ˆ ˆ
= − {d}T  [N ]T {B} dV + [N ]T {T } dS 
Ve Se

If concentrated loads {p} are applied at the element nodes, the work done by them is
{d}T {p} and is added to the total work done by the external forces
The expression inside the parenthesis results in a n × 1 vector known as the consistent
element force vector. This operation can be thought of as replacing the forces acting in
the element by nodal forces which are equivalent to the distributed forces in the sense
that they do the same amount of work. The term consistent means that the equivalency
CHAPTER 2. FINITE ELEMENT FORMULATION OF ELASTIC CONTINUUM 35

is based on the same shape function as that was used in creating the stiffness matrix.
Denoting the consistent force vector by {f } we get
W = − {d}T {f } (2.13)
The potential energy of the system is obtained as

Πp = Π + W
1
= {d}T [k] {d} − {d}T {f } (2.14)
2
Thus the potential energy is expressed in terms of n unknown displacements contained
in{d}. For stable equilibrium the potential energy is minimized with respect to each of
these n unknown displacements, i.e.,
∂Πp
=0
∂ {d}
where the potential energy is minimized with respect to each of the n unknown elements
of {d}. This results in
[k] {d} = {f } (2.15)
This set of n equation relates the nodal displacements and nodal forces of the elements,
and is called as element equations.

2.9 Formulation of system equations


Once the element equations are formulated, they can be assembled to obtain system
equations following the procedure outlined in Section 1.5.4.2. Alternatively, the principle
of minimum potential energy can be applied to the whole system at once. If there are m
elements in total, the potential energy of the whole system is

X
m
Πp = (Πp )i
i=1
Xm
1 Xm
= {d}Ti [k]i {d}i − {d}Ti {f }i
i=1
2 i=1

Let there are N degrees of freedom in the whole system. We can thus expand the element
nodal displacement vector {d}i to include all the degrees of freedom provided that the
stiffness matrix [k]i is expanded accordingly by adding zeros at the columns and rows not
belonging to the element degrees of freedom. The nodal force vector {f }i is expanded
accordingly by adding zeros at those rows which correspond to DOF not contained in the
element. Let the displacement vector thus expanded and containing unknown displace-
ments at all the DOFs be denoted by {D}.!
T
Xm
T
X
m
Πp = {D} [k]i {D} − {D} {f }i
i=1 i=1

The sum of the stiffness matrices of all the elements, each expanded to cover all the DOFs
of the system is the system stiffness matrix, [K] and the sum of the nodal force vector of
all the element, each expanded to cover all the DOFs of the system is the system force
vector {F } defined as
CHAPTER 2. FINITE ELEMENT FORMULATION OF ELASTIC CONTINUUM 36

X
m
[K] = [k]i (2.16)
i=1
Xm
{F } = {f }i (2.17)
i=1

The potential energy can thus be written as


1
Πp = {D}T [K] {D} − {D}T {F }
2
If there are concentrated loads {P } applied at some DOFs, then the work done by them
is {D}T {P }, and the potential energy becomes
1
Πp = {D}T [K] {D} − {D}T {F } − {D}T {P }
2
The minimization of potential energy with respect to each of the N unknown displace-
ments contained in the vector {D} results in a set of N equations which can be written
as
[K] {D} = {F } + {P } (2.18)
which is known as the finite element equation of the system.

2.10 Transformation of coordinate system


The summation of element stiffness matrix and nodal force vector implied in Equations
2.16 and 2.18 assumes that the matrices are formulated in a common coordinate system for
all the elements. In, on the other hand, these matrices are formulated in a local coordinate
system, chosen separately for each element, a proper transformation
 0 of coordinates must
be applied
 0 before the summation is carried out. Let k represent
 0 the element stiffness
matrix d represent the element displacement vector, and f represent the element
force vector, all formulated in localcoordinate system. If the coordinate transformation
0
matrix is denoted by [T ] such that d = [T ] {d}, the transformation of stiffness matrix
and element force vector can be obtained as

h 0i
[k] = [T ]T k [T ]
n 0o
{f } = [T ]−1 f

2.11 Application of displacement boundary conditions


The force boundary conditions are already contained in the system Equation 2.17 in the
sense that the element nodal forces were constructed from the given force boundary con-
ditions. The system stiffness matrix [K] becomes a singular matrix if rigid body motion
is allowed in the system. The rigid body motion can be eliminated from the system equa-
tions by using the knowledge of specified displacements within the system. The degrees
of freedom corresponding to the known displacements are then removed from the system
equations, to obtain a reduced set of equations. Let us assume that out of N unknown
displacements contained in {D}, c number of displacements are known (constrained) and
CHAPTER 2. FINITE ELEMENT FORMULATION OF ELASTIC CONTINUUM 37

these known displacements are collected in a vector {Dc }. The remaining N − c dis-
placements are collected in a separate vector {Du }. The DOF are rearranged, and the
displacement vector is partitioned as
 
{Dc }
{D} = (2.19)
{Du }
The force vector is partitioned similarly  
{Fc }
{F } = (2.20)
{Fu }
The rows and columns of the stiffness matrix [K] needs to be rearranged accordingly, and
it is partitioned as  
[Kcc ] [Kcu ]
[K] = (2.21)
[Kuc ] [Kuu ]
Where the sub matrix [Kcc ] is a c×c matrix formed by taking only those rows and columns
from [K] which correspond to the DOFs of {Dc }. The sub matrix [Kcu ] is a c × N − c
sub matrix formed by taking from [K] those rows which correspond to the DOFs of {Dc }
and those columns which correspond to the DOFs of {Du }. Owing to the symmetry of
[K], we have [Kuc ] = [Kcu ]T . The sub matrix [Kuu ] is a N − c × N − c matrix formed by
taking only those rows and columns from [K] which correspond to the DOFs of {Du }. In
the partitioned form, the
 finite elementequation
 can
 be  writtenas
[Kcc ] [Kcu ] {Dc } {Fc }
=
[Kuc ] [Kuu ] {Du } {Fu }
Performing the matrix multiplication on the left hand side we obtain

[Kcc ] {Dc } + [Kcu ] {Du } = {Fc }


[Kuc ] {Dc } + [Kuu ] {Du } = {Fu }

Where the displacements are specified, the forces are generally unknown, and in many
cases they are the support reactions, as represented by {Fc }. The forces at the rest of
the DOFs as represented by {Fu } should be known. Then from the second of the two
equations above we obtain

[Kuu ] {Du } = {Fu } − [Kuc ] {Dc }


{Du } = [Kuu ]−1 ({Fu } − [Kuc ] {Dc }) (2.22)

If the specified displacements are all zero, i.e., {Dc } is a null vector, then we have
{Du } = [Kuu ]−1 {Fu }. Thus, when the specified displacements are all zero, the reduced
set of equations can simply be obtained by removing the rows and columns of the system
matrices, which correspond to the DOFs with specified zero displacements. In the general
case, when non-zero displacements are specified at certain DOFs, Equation 2.22 should
be used. One practical example when such cases might arise is when supports, which
are otherwise expected to be fixed, move due to some imperfections or other mechanical
causes. Once the unknown displacements are obtained, the unknown forces are given by
the
{Fc } = [Kcc ] {Dc } + [Kcu ] {Du } (2.23)
CHAPTER 2. FINITE ELEMENT FORMULATION OF ELASTIC CONTINUUM 38

2.12 Computing the strain and stress field


The strain field within each each element can be computed from its nodal displacements.
From the vector {D} we extract only those elements which correspond to the DOFs of the
element. This provides us the element nodal displacement vector {d}. The strain field is
then obtained by using the strain matrix

{e} = [S] {d}


and the stress field is obtained by using the elasticity matrix

{σ} = [E] [S] {d}

2.13 Summary
We have formally derived the finite element form of equilibrium equations of a elastic
continuum. A continuum is divided into a finite number of smaller parts called elements.
The elements share common nodes, and therefore displacement continuity among elements
at the nodes is guaranteed. The displacement field within each element is interpolated
in terms of the nodal displacements by using shape functions. Thus a continuous dis-
placement field is approximated by a finite number of unknown nodal displacements. By
minimizing the potential energy of the system with respect to each of the unknown dis-
placements, we can establish a system of linear equations relating the nodal forces and
displacement with the elastic and geometric properties of the continuum. Such a set of
linear equation is often called as the finite element approximation of equilibrium equa-
tions. The finite element equations, after applying proper boundary conditions, can be
solved to obtain nodal displacements. Nodal displacements are then interpolated to ob-
tain displacement field, and subsequently strain and stress fields. This chapter developed
the procedure in a general three dimensional setting. In the subsequent chapters, we
apply this formulation to certain examples commonly encountered in solid and structural
mechanics.
Chapter 3

Development of truss members

3.1 Introduction
In this chapter, we illustrate the finite element formulation of Chapter 2 for a special
type of structural member known as a truss element. We derive the stiffness matrix of
truss element using a local coordinate system, chosen with the element in mind. The
coordinate transformation equations applicable to truss elements arranged in a plane as
well as a three dimensional space are developed. Solved numerical examples are presented
for illustration.

3.2 Truss element


A truss element is a prismatic (constant cross-sectional area) bar that can deform either
in tension or in compression. Following assumptions are implicit in formulating a truss
element.
1. The bar cannot sustain shear forces or bending moments.
2. Only longitudinal (along the axis) displacement is considered, transverse displace-
ment is ignored.
3. The material of the bar is linearly elastic.
4. Loads are only applied at the joints between different members; no intermediate
loads are applied. This assumption is necessary to ensure that the member is not
exposed to shear forces and bending moments.

3.3 Shape function


Consider a prismatic bar of cross sectional area A and length L as shown in Figure
3.1. A local coordinate system xy is selected as shown in the figure. Only longitudinal
displacements are considered, and u(x, y) describes the displacement field. Any cross
section of the bar undergoes constant displacement, the dependance of u on y is removed.
Thus the displacement field can be defined as u(x) and the problem is one dimensional.
We select two nodes at the ends of the bar as indicated with the red dots. The axial
displacements of the two nodes are denoted as d1 and d2 . It is required to interpolated
the displacement field u(x) within the bar from the nodal displacements d1 and d2 .

39
CHAPTER 3. DEVELOPMENT OF TRUSS MEMBERS 40

u
x,
d2
2
y

Y L
d1 1

X
Figure 3.1: A truss element

Polynomial functions provide a suitable interpolation model in many finite element appli-
cations, as it is quite easy to perform differentiation and integration of such functions. In
this case, the displacement field is to be interpolated in terms of two quantities, d1 and
d2 , and thus a linear interpolation is appropriate. Let us consider a linear function
u(x) = a1 + a2 x (3.1)
Using the condition u(0) = d1 , we get , a1 = d1 . In addition, using u(L) = d2 , we obtain
a2 = (d2 − d1 ) /L . Thus the displacement field is given by
d2 − d1
u = d1 + x
 L x
x
= d1 1 − + d2
L  L
  d1
= 1 − Lx Lx
d2
n 0o
= [N ] d

where
 0 [N ] is the shape function matrix which operates on the nodal displacement vector
d to give the displacement field. The columns of the shape function matrix N1 = 1 − Lx
and N2 = Lx are called the shape functions. Note that N1 is 1 at node 1, and 0 and node
2; while N2 is 1 at node 2, and 0 at node 1.

3.4 Stiffness matrix


We now consider the kinematic behavior of the truss element. Since this is a one dimen-
sional problem, there is only one component ofstrain
 relevant to this problem.
∂u ∂
{exx } = = {u}
∂x ∂x
∂
Comparing this with Equation 2.2 we see that the differential operator matrix is [L] = ∂x
and has only one element for this one dimensional problem. The stress strain relation for
a linear elastic behavior is
σxx = Eexx
CHAPTER 3. DEVELOPMENT OF TRUSS MEMBERS 41

Comparing this with Equation 2.4, we observe that the elasticity matrix [E] = E has only
one element in this case. The strain matrix is obtained as
∂   1 
[S] = [L] [N ] = 1 − Lx Lx = 1 −1
∂x L
Finally the stiffness matrix is obtained as
h 0i ˆ
k = [S]T [E] [S] dV
V
ˆL  
A 1  
= E 1 −1 dx
L2 −1
0
  ˆL
AE 1 −1
= dx
L2 −1 1
 0
AE 1 −1
=
L −1 1

3.5 Plane truss : coordinate transformation


A plane truss is a structures made from an assembly of truss elements arranged in a plane
such as the one shown in Figure 1.10. In the element coordinate system, each node of
a truss member has one DOF. For a truss member oriented arbitrarily in a 2D plane,
each node has 2 DOFs, which are displacements parallel to the coordinate axes. Let us
consider the truss element of Figure 3.2 which is inclined to the global X axis at an angle
of α.
u
D4 x,
d2 α
2
y D2 D3

D1 L
Y
d1 1

X
Figure 3.2: A truss element: local and global DOFs

The nodal displacement vector in the local coordinate system is


 ’  T
d = d1 d2
and that in the global coordinate system
 is 
{d} = D1 D2 D3 D4
CHAPTER 3. DEVELOPMENT OF TRUSS MEMBERS 42

Let the direction cosines of the local x axis be defined as lx = cos α and mx = sin α .
Using geometry we get

d1 = D1 lx + D2 mx
d2 = D3 lx + D4 mx

Writing in matrix form we have


 
   
 D1 

 
d1 lx mx 0 0 D2
=
d2 0 0 lx mx 
 D3 

 
D4
 ’
d = [T ] {d} (3.2)

where [T ] is the transformation


 0 matrix from global to local coordinates. The inverse
transformation is {d} = [T ]−1 d . It can be shown that the inverse of [T ] is equal to its
transpose. Therefore we have n 0o
{d} = [T ]T d (3.3)
See Section 1.5.4.2 for a numerical example of solving a plane truss.

3.6 Space truss : coordinate transformation


If the truss members are arranged in a 3D space, the resulting structure is called a space
truss. Each node of a space truss has 3 DOFs, corresponding to the displacement parallel
to the 3 coordinate axes. In this section we develop the coordinate transformation matrix
for truss elements in a 3D space. Consider a truss element in a 3D space. The coordinates
of the nodes 1 and 2 with respect to a right handed Cartesian frame XYZ are (X1 , Y1 , Z1 )
and (X2 , Y2 , Z2 ) respectively. The DOFs in the local coordinate system are d1 and d2 and
those in the global coordinate system are D1 ..D6 . We seek a transformation between the
nodal displacements in the two coordinate systems. Let us define the direction cosines of
the local x axis with respect to the global X, Y, and Z axis as lx , mx , and nx , respectively.
Note that lx , mx , and nx are the cosine of the angles between local x axis and global X ,
Y, and Z axes, respectively. The direction cosines and the length of the member can be
easily obtained from the coordinates of the nodes as follows.

q
L = (X22 − X12 ) + (Y22 − Y12 ) + (Z22 − Z12 ) (3.4)
X 2 − X1
lx = (3.5)
L
Y2 − Y1
mx = (3.6)
L
Z2 − Z1
nx = (3.7)
L
From geometry we have the following relation between the global and local displacements.

d1 = lx D1 + mx D2 + nx D3
d2 = lx D4 + mx D5 + nx D6
CHAPTER 3. DEVELOPMENT OF TRUSS MEMBERS 43

These equations can be written in matrix form as


 

 D1 


 

   
 D2 

 
d1 lx mx nx 0 0 0 D3
=
d2 0 0 0 lx mx nx 
 D4 


 


 D5 

 
D6
n 0o
d = [T ] {d}

with [T ] as the transformation matrix. The inverse relation can be written as {d} =
0
[T ]T d .

D5
y d2 x, u
D2
L D4
2
d1 1
Y D6
D1
D3
z
X

Figure 3.3: A truss element in 3D space: local and global DOFs

3.6.1 Solved example


Consider a truss structure consisting of 3 members as shown in Figure 3.4. The four nodes
of the truss are indicated with red dots. The nodes are numbered from 1 to 4, and their
coordinates with respect to the global frame XYZ are shown in the figure. Nodes 2, 3,
and 4 are hinge-supported, which means that truss members at these nodes can freely
rotate, but cannot translate (move) in any direction. At node 1, the translation in the Z
direction is constrained, whereas the node is free to move in X and Y directions. A load
of 1kN is applied at node 1 as shown in the figure. All the members are made up of a
material having E = 200GP a and have a cross sectional area of 0.01m2 . It is required to
find the displaced configuration of the truss.
CHAPTER 3. DEVELOPMENT OF TRUSS MEMBERS 44

3 (0,2,-1)

0,-1)
4 (0,

Z 1 1 (2,0,0)

2 (0,-1,0) X
100 kN
Figure 3.4: A space truss

The numbering of members is as indicated in Figure 3.4. Each node has 3 DOFs,
which are displacements parallel to the X, Y, and Z directions. The degrees of freedom
are assigned starting at the first node, where X, Y, and Z displacements are numbered as
D1 , D2 , and D3 , respectively. DOFs at other nodes are numbered accordingly. According
to this convention, the DOFs at a node i are D3i−2 , D3i−1 , D3i , corresponding to the X,
Y, and Z directions, respectively. The geometrical properties of the members are shown
in Table 3.1.

Table 3.1: Gemological properties of the members


Member Node 1 Node 2 L lx mx nx DOFs
1 1 2 2.34 -0.89 -0.45 0.00 1,2,3,4,5,6
2 1 3 3.00 -0.67 0.67 -0.33 1,2,3,7,8,9
3 1 4 2.34 -0.89 0.00 -0.45 1,2,3,10,11,12

The element stiffness matrices are computed next.

Element 1
h 0 i  8.94 −8.94 
k = × 108
−8.94 8.94
CHAPTER 3. DEVELOPMENT OF TRUSS MEMBERS 45

 
−0.89 −0.45 0 0 0 0
[T ] =
0 0 0 −0.89 −0.45 0
 
7.15 3.58 0 −7.15 −3.58 0
 3.58 1.79 0 −3.58 −1.78 0 
h 0i  
 0 0 0 0 0 0 
[k] = [T ]T k [T ] = 
 −7.15 −3.57 0

 7.15 3.58 0 
 −3.58 −1.79 0 3.58 1.79 0 
0 0 0 0 0 0

When this matrix is added to DOFs 1, 2, 3, 4, 5, and 6, the system stiffness matrix is
updated as
 
7.15 3.58 0 −7.15 −3.58 0 0 0 0 0 0 0
 3.58 1.79 0 −3.58 −1.78 0 0 0 0 0 0 0 
 
 0 0 0 0 0 0 0 0 0 0 0 0 
 
 −7.15 −3.57 0 7.15 3.58 0 0 0 0 0 0 0 
 
 −3.58 −1.79 0 3.58 1.79 0 0 0 0 0 0 0 
 
 0 0 0 0 0 0 0 0 0 0 0 0 
[K] = 
 0


 0 0 0 0 0 0 0 0 0 0 0 
 0 0 0 0 0 0 0 0 0 0 0 0 
 
 0 0 0 0 0 0 0 0 0 0 0 0 
 
 0 0 0 0 0 0 0 0 0 0 0 0 
 
 0 0 0 0 0 0 0 0 0 0 0 0 
0 0 0 0 0 0 0 0 0 0 0 0

Element 2
h 0 i  6.67 −6.67 
k = × 108
−6.67 6.67

 
−0.67 0.67 −0.33 0 0 0
[T ] =
0 0 0 −0.67 0.67 −0.33
 
2.96 −2.96 1.48 −2.96 2.96 −1.48
 −2.96 2.96 −1.48 2.96 −2.96 1.48 
h 0i  
 1.48 −1.48 0.74 −1.48 1.48 −0.74 
[k] = [T ] k [T ] = 
T 
 −2.96 2.96 −1.48 2.96 −2.96 1.48  × 10
8

 
 2.96 −2.96 1.48 −2.96 2.96 −1.48 
1.48 1.48 −0.74 1.48 −1.48 0.74

When this matrix is added to DOFs 1, 2, 3, 7, 8, and 9, the system stiffness matrix is
updated as
CHAPTER 3. DEVELOPMENT OF TRUSS MEMBERS 46

 
10.11 0.61 1.48 −7.15 −3.58 0 −2.96 2.96 0 0 0 0
 0.61 4.75 −1.48 −3.58 −1.78 0 2.96 −2.96 0 0 0 0 
 
 1.48 −1.48 0.74 0 0 0 −1.48 1.48 0 0 0 0 
 
 −7.15 −3.57 0 7.15 3.58 0 0 0 0 0 0 0 
 
 −3.58 −1.79 0 3.58 1.79 0 0 0 0 0 0 0 
 
 0 0 0 0 0 0 0 0 0 0 0 0 
[K] = 

 × 108

 −2.96 2.96 −1.48 0 0 0 2.96 −2.96 0 0 0 0 
 2.96 −2.96 1.48 0 0 0 −2.96 2.96 0 0 0 0 
 
 −1.48 1.48 −0.74 0 0 0 1.48 −1.48 0 0 0 0 
 
 0 0 0 0 0 0 0 0 0 0 0 0 
 
 0 0 0 0 0 0 0 0 0 0 0 0 
0 0 0 0 0 0 0 0 0 0 0 0

Element 3
h 0 i  8.94 −8.94 
k = × 108
−8.94 8.94

 
−0.89 0 −0.44 0 0 0
[T ] =
0 0 0 −0.89 0 −0.44
 
7.15 0 3.58 −7.15 0 −3.58
 0 0 0 0 0 0 
h 0i  
 3.58 0 1.79 −3.58 0 −1.79 
[k] = [T ]T k [T ] = 
 −7.15 0
 × 108
 −3.58 7.15 0 3.58 
 0 0 0 0 0 0 
−3.58 0 −1.79 3.58 0 1.79

When this matrix is added to DOFs 1, 2, 3, 10, 11, and 12, the system stiffness matrix is
updated as
 
17.27 0.62 5.06 −7.15 −3.58 0 −2.96 2.96 −1.48 −7.15 0 −3.58
 0.62 4.75 −1.48 −3.58 −1.78 0 2.96 −2.96 1.48 0 0 0 
 
 5.06 −1.48 2.53 0 0 0 −1.48 1.48 −0.74 −3.58 0 −1.79 
 
 −7.15 −3.58 0 7.15 3.58 0 0 0 0 0 0 0 
 
 −3.58 −1.79 0 3.58 1.79 0 0 0 0 0 0 0 
 
 0 0 0 0 0 0 0 0 0 0 0 0 
[K] = 

 × 108

 −2.96 2.96 −1.48 0 0 0 2.96 −2.96 1.48 0 0 0 
 2.96 −2.96 1.48 0 0 0 −2.96 2.96 −1.48 0 0 0 
 
 −1.48 1.48 −0.74 0 0 0 1.48 −1.48 0.74 0 0 0 
 
 −7.15 0 −3.57 0 0 0 0 0 0 7.15 0 3.58 
 
 0 0 0 0 0 0 0 0 0 0 0 0 
−3.58 0 −1.79 0 0 0 0 0 0 3.58 0 1.79

The force vector consists of 12 elements, of which there are 1 reaction force at node 1,
and 3 reaction forces at nodes 2, 3, and 4 each. At node 1, F2 = −100000N , and F1 = 0.
The force vector is thus
 
{F } = F1 F2 F3 F4 F5 F6 F7 F8 F9 F10 F11 F12
 
= 0 −1000 F3 F4 F5 F6 F7 F8 F9 F10 F11 F12
CHAPTER 3. DEVELOPMENT OF TRUSS MEMBERS 47

The displacement boundary conditions is that D3 = D4 = D5 = D6 = D7 = D8 = D9 =


D10 = D11 = D12 = 0, i.e., DOFs, 3, 4, 5, 6, 7, 8, 9, 10, 11, and 12 are constrained. We
remove these rows and columns from the stiffness matrix to obtain a reduced stiffness
matrix.  
17.27 0.62
[Kuu ] = × 108
0.62 4.75
The reduced force vector is    
F1 0
{Fu } = = × 105
F2 −1
The unknown displacement vector is given by
 
D1
{Du } = = [Kuu ]−1 {Fu }
D2
 
0.0075
= mm
−0.2114

Finally the support reactions can be obtained as


 T
{Fc } = F3 F4 F5 F6 F7 F8 F9 F10 F11 F12
= [Kcu ] {Du }
 
5.06 −1.48
 −7.15 −3.57 
 
 −3.58 −1.78 
 
 0 0 
  
 −2.96 2.96  0.0075
  5
= 
2.96 −2.96  −0.2114 × 10
 
 −1.48 1.48 
 
 −7.15 0 
 
 0 0 
−3.58 0
 T
= 35.13 70.26 35.13 0 −64.87 64.87 −32.44 −5.39 0 −2.69 kN

The support reactions are shown in Figure 3.5 with green arrows, while the applied load
is shown with a black arrow. Note that the results satisfy equilibrium
X
FX = 70.26 − 64.87 − 5.39 = 0
X
FY = −100 + 35.13 + 64.87 = 0
X
FZ = 35.13 − 32.44 − 2.69 = 0
CHAPTER 3. DEVELOPMENT OF TRUSS MEMBERS 48

Y
64.87

64.87
32.44

2.69
4
5.39

35.13 35.13
Z
2

70.26 X
100

Figure 3.5: Applied forces and support reactions of the space truce of Figure 3.4.

The stresses in the members can now be obtained by using the nodal displacements. For
example, let us consider member 1. The DOFs associated with the member are 1,2,3,4,5,
and 6. Hence the nodal displacement vector for the element in global coordinate system
is
 T
{d} = D1 D2 D3 D4 D5 D6
 T
= 0.0075 −0.2114 0 0 0 0 × 10−3

The nodal displacement vector in the element coordinate system is


n 0o
d = [T ] {d}
 

 0.0075 


 

  −0.2114 

 
−0.89 −0.45 0 0 0 0 0
= × 10−3
0 0 0 −0.89 −0.45 0   0 


 


 0 

 
0
 
0.88
= × 10−3
0

The strain in the element coordinate system is given by


CHAPTER 3. DEVELOPMENT OF TRUSS MEMBERS 49

n 0o
{e} = {exx } = [S] d
 
1  0.88
= 1 −1 × 10−3
L 0
= 3.93 × 10−5

Finally the stress is given by

{σ} = [E] {e}


⇒ σxx = 3.93 × 10−5 E
= 7.85 × 106 Pa
Chapter 4

Development of beam members

4.1 Introduction
Beams are the most common type of structural component in Civil and Mechanical en-
gineering. This chapter introduces the basic mechanics of beams along with their finite
element formulations. A brief description of terminologies related to the mathematical
modeling of beams is followed by engineering modeling assumptions, and mechanical sim-
plifications. Finite element formulation of beams is presented next.

4.2 Beam element


A beam is a bar-like structural component which is primarily required to support trans-
verse loads and carry them to the supports. A bar-like element has one of its dimensions
considerably larger than the other two. This dimension is called the longitudinal dimen-
sion or beam axis. The planes normal to the beam axis are called as cross sections. A
plane passing through the beam axis is called as the longitudinal plane.
Beams are able to resist transverse loads mainly through bending. Bending action
induces compressive longitudinal strains in one side of the beam and tensile strains in the
other. These regions of tension and compression are separated by a neutral (no strain)
surface called as the neutral surface (see Figure 4.1). The combination of tensile and
compressive stresses produces internal bending moment, which is the primary mechanism
by which loads are transferred to the supports.

neutral surface
compression

tension

Figure 4.1: Transverse loads are resisted by bending action.

We consider beams have a straight longitudinal axis and are prismatic (cross section
is constant).

50
CHAPTER 4. DEVELOPMENT OF BEAM MEMBERS 51

4.2.1 Mathematical model: Bernoulli-Euler beam


One-dimensional models of beams are constructed on the basis of beam theories. Because
beams are actually three-dimensional bodies, their simplification as a one-dimensional
model involves some for of approximation of the underlying mechanics. The simplest of
such models is based on the Bernoulli-Euler beam theory, also called as classical beam
theory. Following assumptions are included in the classical beam theory.

Assumptions of classical beam theory


1. The longitudinal axis is straight and the cross section of the beam has a longitudinal
plane of symmetry.

2. The cross section is constant.

3. Plane sections originally normal to the longitudinal axis of the beam remain plane
and normal to the deformed longitudinal axis upon bending.

4. The internal strain energy of the beam is mostly contributed by bending stresses.

5. The material of the beam is assumed to be elastic, isotropic, and homogeneous.

4.2.1.1 Element coordinate system


Under transverse loading, one of the extreme surfaces shortens, while the other elongates.
Therefore a neutral surface that undergoes no axial strain exists between the extreme
surfaces. For a homogeneous cross section, the neutral surface lies on centroid of the cross
section. The intersection of the neutral surface with a cross section defines the neutral
axis at that cross section. The Cartesian axes for analysis of a plane beam are selected
as follows

• Axis x lies along the longitudinal beam axis, at neutral axis height.

• Axis y lies in the symmetry plane of the cross section and points upwards.

• Axis z is directed along the neutral axis, forming a right handed coordinate system
with x and y. The origin is placed at the leftmost section. The total length (or
span) of the beam is denoted as L.

y
y, uy
qy(x)
neutral surface

x, ux z neutral axis

centroid

symmetry plane
L

Figure 4.2: Local coordinate system of a plane beam.


CHAPTER 4. DEVELOPMENT OF BEAM MEMBERS 52

4.2.1.2 Kinematics
The motion of the plane beam in the xy plane is described by the two dimensional
displacement field
 
ux (x, y)
{u(x, y)} =
uy (x, y)
where ux (x, y) and uy (x, y) are the axial and transverse displacement components, re-
spectively, of an arbitrary point on the beam. The motion in the z direction which is
primarily due to the Poisson’s ratio effects, is neglected. The assumption of normality
of cross section allows us to simplify the displacement field, reducing it to a one dimen-
sional problem. Consider a small section of a beam as shown in Figure 4.3. We focus
our attention on a cross section where two points A , and B are marked. After bending,
these points are denoted as A’, and B’. The neutral surface has rotated counter clockwise
(CCW) by an angle of θ. Point A lies on the neutral surface, and therefore does not
move in the x direction. The displacement of A in the y direction is uy (x, 0). For small
displacements, and because plane sections normal to the neutral surface remain plane and
normal to the neutral surface after bending, the displacement of B in the y direction is
the same as that of A . Hence we can say that at any given x, uy is independent of y, and
can be written as uy (x). The displacement of B in the x direction is denoted as ux (x, y),
and can be obtained as

ux (x, y) = −y sin θ (4.1)

ux(x,y)
θ
θ
B'
y

A'

uy(x,y = 0)

B
y

Figure 4.3: Kinematics of a beam.

For small displacements θ ≈ sin θ , therefore ux (x, y) = −yθ. The angle of rotation θ is
the slope of the deflection curve (neutral surface after bending) and can be represented
as the derivative of deflection, i.e., θ(x) ≈ tan θ(x) = ∂uy (x)/∂x . Hence we obtain
∂uy (x)
ux = −y
∂x
CHAPTER 4. DEVELOPMENT OF BEAM MEMBERS 53

This shows that the x component of displacement can be determined once the y component
is known. Finally the displacement field is written
 as
{u(x)} = uy (x) (4.2)
The longitudinal strain is obtained as
∂ux ∂ 2 uy (x)
exx = = −y (4.3)
∂x ∂x2

4.2.1.3 Shape function


We have seen from the previous section, that the displacement field is described by the
traverse displacement and the rotation of the neutral surface of the beam. Therefore we
select the transverse displacement and rotation as the degrees of freedom at the nodes
of a beam element, as shown in Figure. The rotational degrees of freedom d2 and d4 are
considered positive if they are CCW.
y

d1 d3

d4
d2
x
1 2

Figure 4.4: Local coordinate system of a plane beam.

With the 4 DOF available, a cubic polynomial can be used to interpolate the displacement
field within the element, i.e.,

uy (x) = a + bx + cx2 + dx3 (4.4)


Differentiating the displacement field gives us the rotation field, i.e,
θ(x) = b + 2cx + 3dx2 (4.5)
The constants a, b, c, d, can be obtained by using the displacement DOFs. Taking x = 0,
and uy = d1 , θ = d2 we get
d1 = a + b(0) + c(0) + d(0) (4.6)
d2 = a(0) + b + c(0) + d(0) (4.7)
Similarly taking x = L, and uy = d3 ,θ = d4 we obtain
d3 = a + b(L) + c(L2 ) + d(L3 ) (4.8)
d4 = a(0) + b + c(2L) + d(3L2 ) (4.9)
Combining Equations 4.6to 4.9 we get
    

 d1 
 1 0 0 0 
 a 

     
d2 0 1 0 0  b
= 2 3  (4.10)

 d3 
 1 L L L 
 c 

   
d4 0 1 2L 3L2 d
Inverting the equation, we can determine the constants of interpolations as
CHAPTER 4. DEVELOPMENT OF BEAM MEMBERS 54

   −1  

 a  1 0 0 0  d1 
    

 

b 0 1 0 0  d2
=
 (4.11)

 c 
 1 L L2 L3    d3 

   
d 0 1 2L 3L2 d4
And from Equation 4.4 we obtain the displacement field

uy (x) = a + bx + cx2 + dx3


 

 a 

   
2 3 b
= 1 x x x

 c 

 
d
 −1  
1 0 0 0 
 d1 

  
  0 1 0 0  d2
= 1 x x2 x3  1

L L 2 L3    d3 

 
0 1 2L 3L2 d4
 
 d1 
h         i d 
 
2x3 3x2 2x2 x3 3x2 2x3 x3 x2
= 1+ L3
− L2
x− + L2 L2
− L3 L2
− L 2
(4.12)
L 
 d3 

 
d4
= [N (x)] {d} (4.13)

where [N (x)] is the required shape function matrix. The four columns of [N (x)] are called
as the shape functions. The four columns of the shape function matrix [N (x)] are denoted
as N1 (x), N2 (x), N3 (x), and N4 (x) . These functions represent the deflected shape of the
neutral surface when the corresponding DOF is moved by an unit amount, and all other
DOFs are fixed. For example, N1 (x) represents deflected shape of the beam neutral surface
for d1 = 1, and d2 = d3 = d4 = 0. These functions thus act to represent the deflected
shape of the beam, and when scaled with the displacement at their corresponding DOF,
provide the actual deflection of the beam. This can also be understood from Figure 4.5,
where the shape functions are plotted as a function of x , along the length of the beam.
Note that any shape function Ni is equal to 1 at the DOF i , and at other DOFs.

d2=0 N1(x) N2(x)

d1=1 d2=1
d3=d4=0 d3=d4=0

d1=0

N3(x) d4=0 N4(x)

d3=1
d1=d2=0 d1=d2=0 d3=0

d4=1

Figure 4.5: Shape functions of a beam element.


CHAPTER 4. DEVELOPMENT OF BEAM MEMBERS 55

4.2.1.4 Strain matrix


With the displacement field obtained from the nodal displacements and the shape function
matrix, we can now determine the shape function matrix. To do this, we start from
the definition of longitudinal strain (see Equation 4.2), from where it is clear that the
differential operator matrix is

∂2
[L] = −y
∂x2
The strain matrix is obtained as

[S] = [L] [N ]
∂2  
= −y 2 N1 N2 N3 N4
∂x
∂2 00
For notational simplicity we denote N
∂x2 i
as Ni , and write the strain matrix as
 00 00 00 00 
[S] = −y N1 N2 N3 N4 (4.14)

4.2.1.5 Elasticity matrix


The relation between longitudinal stress and strain is σxx = Eexx , where E is the Young’s
modulus of elasticity. This shows that the elasticity matrix for a classical beam formula-
tion contains only one element, E

[E] = E (4.15)

4.2.1.6 Element stiffness matrix


The stiffness matrix of a beam element with cross sectional area A and length L can now
be obtained using the procedure described in Section 2.8 as

ˆ
[k] = [S]T [E] [S] dV
V
ˆ
= E [S]T [S] dV
V
 00 00 00 00 00 00 00 00 

ˆL N100 N100 N1 N2 N1 N3 N1 N4 ˆ
 N2 N1 00 00
N2 N2
00 00
N2 N3 N2 N4 
00 00

= E  N300 N100
 y 2 dA dx (4.16)
N3 N4 
00 00 00 00 00 00
N3 N2 N3 N3
0 00 00 00 00 00 00 00 00
N4 N1 N4 N2 N4 N3 N4 N4 A
´
In the above equation the integral A y 2 dA is the second moment of area of the beam
about the z axis (neutral axis), sometimes called as second moment of inertia or simply
moment of inertia of the beam, and is denoted as Izz . For a rectangular beam with
thickness t and height h, the second moment of area is given by

th3
Izz =
12
CHAPTER 4. DEVELOPMENT OF BEAM MEMBERS 56

Replacing for this quantity in Equation 4.16, the element stiffness matrix is obtained as
 00 00 00 00 00 00 00 00 

ˆ L N1 N1 N1 N2 N1 N3 N1 N4
 N200 N100 N200 N200 N2 N3 N2 N4 
00 00 00 00

[k] = EIzz   N300 N100 N300 N200


 dx
N3 N3 N3 N4 
00 00 00 00

0 00 00 00 00 00 00 00 00
N4 N1 N4 N2 N4 N3 N4 N4
The elements of the stiffness matrix can be easily evaluated by performing the integration
operations. One example is shown below

ˆL
00 00
k11 = EIzz N1 N1 dx
0
ˆL  2
12x 6
= EIzz 3
− 2 dx
L L
0
  ˆL  
36 4x2 4x
= EIzz − +1 dx
L4 L2 L
0
  
36 4L
= EIzz − 2L + L
L4 3
12EIzz
=
L3
The other elements of the stiffness matrix are determined similarly, and the resulting
stiffness matrix is given below
 
12 6L −12 6L
EIzz  6L 4L2 −6L 2L2 
[k] = 3  
L  −12 −6L 12 −6L 
6L 2L2 −6L 4L2

4.2.1.7 Consistent force vector


The consistent force vector for a beam element can be determined by using the procedure
discussed in Section 2.8. Let us consider a beam element as shown in Figure 4.2. The
body forces are neglected, and the traction forces are applied on the top surface of the
beam as shown in the figure. The traction force is qy (x) force per unit length of the beam.
If t is the thickness of the beam, the traction measured as force per unit surface area is
qy (x)/t. For simplicity the traction is considered to be uniform through the length of the
beam and is therefore denoted as qy /t. Based on the developments from Section 2.8, the
consistent load vector is obtained as
CHAPTER 4. DEVELOPMENT OF BEAM MEMBERS 57

ˆ
{f } = [N ]T {T } dS
Se
ˆ
qy
= [N ]T t dx
t
L
 3

 3x2 
 1 + 2x3 − L2 
ˆ 
L  L
2 3 

x − 2xL + Lx 2
= qy 2 3 dx
 3x
 − 2x 

0  L2
x3
L3
2 

L2
− xL
 L 

 
 L22  
= qy 12
L

 

 −L2
2 
12
Chapter 5

FE formulation of dynamic systems

5.1 Introduction
Dynamic systems are those whose displacements, velocities, strains, stresses and applied
loads are all time-dependent. In deriving the finite element equations of motion, time
derivatives of displacements have to be considered. The first time derivative of displace-
ment is called as velocity. Velocity induces kinetic energy in an inertial system, just as
displacement induces strain energy in an elastic system. Formulation of kinetic energy is
therefore a central concept in deriving finite element equations of dynamic elastic contin-
uum. In this chapter, kinetic energy of an elastic continuum is expressed as a function of
nodal velocities. By using displacement interpolation functions, strain energy of an elastic
continuum was expressed as a function of nodal displacement vector and stiffness matrix
in Chapter 2. In a similar manner, kinetic energy can be expressed as a function of nodal
velocities and mass matrix. By combining the potential energy derived in Chapter 2 with
the kinetic energy, an expression of the Lagrangian functional is obtained. Application of
Hamilton’s principle to the Lagrangian functional leads to the derivation of equilibrium
equations in the finite element formulation of dynamic systems. Finally mass matrices
for some common structural members such as truss, beams, and frames are derived.

5.2 Terminology
Much of the terminology used in this chapter is similar to that in Chapter 2, with the
difference that the field quantities are now a position of both space and time. The dis-


placement field is denoted as U (X, Y, Z, t) where t denotes time. The displacement field
of an element is denoted as →

u (x, y, z, t), where the spatial coordinates (x, y, z) are suitably
selected for the element and can be different from the global spatial coordinates (X, Y, Z).
In a general three dimensional problem the element displacement field consists of three
components, all time-dependent.
 
 ux (x, y, z, t) 


u (x, y, z, t) = uy (x, y, z, t)
 
uz (x, y, z, t)
Using displacement interpolation (shape function) the continuous displacement field can
be expressed in terms of a discrete element displacement vector

−u (x, y, z, t) = [N (x, y, z)] {d(t)}

58
CHAPTER 5. FE FORMULATION OF DYNAMIC SYSTEMS 59

Note that the shape function matrix N (x, y, z) is an interpolation operator in space, and is
independent of time. On the other hand, the nodal displacement vector is time dependent.
To simplify the notations, the dependence on spatial coordinates is considered understood
and not dropped from the notation. For example, displacement field is written as → −u (t).
With this notation the interpolation equation becomes


u (t) = [N ] {d(t)} (5.1)
The time derivative of displacement field gives the velocity field. By differentiating Equa-
tion 5.1 with respect to time, the velocity field is obtained as
∂ n o

−̇ ˙
u (t) = [N ] {d(t)} = [N ] d(t) (5.2)
n o ∂t
˙
where d(t) represents the nodal velocity vector. Time derivative of velocity field gives
the acceleration field written as

−̈ ∂ n˙ o n
¨
o
u (t) = [N ] d(t) = [N ] d(t) (5.3)
n o ∂t
¨
where d(t) is the nodal acceleration vector.

5.3 Kinetic energy


If we consider a small volume of the element dV and if ρ is the mass density of the
material, the mass of the small volume is ρdV . If the velocity at the center of the small
volume is →−̇
u , the kinetic energy of that small volume is
1 T
dψ = ρdV → u →
−̇ −̇
u
2
Using Equation 5.2, and integrating over the volume of the element gives the total kinetic
energy of that element
ˆ n oT n o
1
ψ = ρ d˙ [N ]T [N ] d˙ dV
2
Ve
 
n o ˆ n o
1 ˙  T
= d ρ [N ]T [N ] dV  d˙ (5.4)
2
Ve

The quantity inside the parenthesis in Equation 5.4 is called as the mass matrix of the
element and is denoted as [m]. Thus the kinetic energy can be written as
1 n ˙oT n o
ψ= d [m] d˙ (5.5)
2

5.4 Lagrangian functional


The Lagrangian functional is defined as kinetic energy minus the potential energy. For
the element under consideration, this functional can be written as

l = ψ − πp (5.6)
Using the expression for potential energy from Section 2.8, the Lagrangian can be written
as
1 n ˙oT n o 1
l= d [m] d˙ − {d}T [k] {d} + {d}T {f } (5.7)
2 2
CHAPTER 5. FE FORMULATION OF DYNAMIC SYSTEMS 60

5.5 Hamilton’s principle: Element equations


Hamilton’s principle can be stated as follows
Of all possible time histories of displacement which satisfy the compatibility
equations, boundary conditions, and initial and final conditions at time t1
and t2 , the history corresponding to the actual solution makes the Lagrangian
functional a minimum.
Mathematically, this principle can be expressed as
 t 
ˆ2
δ  ldt = 0
t1

where δ is the variational operator.


Using Equation5.7 we obtain 
ˆt2  n oT n o 
1 ˙ 1
δ d [m] d˙ − {d}T [k] {d} + {d}T {f } dt = 0 (5.8)
2 2
t1

The variation and integration operations can be interchanged. The three different terms
in the above equation are evaluated
 next. 
1
δ {d}T [k] {d} = δ {d}T [k] {d} (5.9)
2
To explicitly illustrate the mathematical operation in Equation 5.9 consider a simple
example of a 2 DOF system.
   T   !
1 1 d k k d
δ {d}T [k] {d} = δ 1 11 12 1
2 2 d2 k12 k22 d2
1 
= δ k11 d21 + 2k12 d1 d2 + k22 d22
2 
1 ∂ 2 2

= k11 d1 + 2k12 d1 d2 + k22 d2 δd1
2 ∂d1
 
1 ∂ 2 2

+ k11 d1 + 2k12 d1 d2 + k22 d2 δd2
2 ∂d2
1
= [(2k11 d1 + 2k12 d2 ) δd1 + (2k12 d1 + 2k22 d2 ) δd2 ]
2  
  k11 d1 + k12 d2
= δd1 δd2
k12 d1 + k22 d2
  
T k11 k12 d1
= δ {d}
k12 k22 d2
= δ {d}T [k] {d}

Similarly we have  n o
1 ˙ T n o n oT n o
δ d [m] d˙ = δ d˙ [m] d˙ (5.10)
2
and  
δ {d}T {f } = δ {d}T {f } (5.11)
CHAPTER 5. FE FORMULATION OF DYNAMIC SYSTEMS 61

Using equations 5.9, 5.10, and 5.11 in equation 5.8 results in


ˆt2 n oT n o ˆt2 ˆt2
δ d˙ [m] d˙ dt − δ {d} [k] {d} dt + δ {d}T {f } dt = 0
T
(5.12)
t1 t1 t1

Variation and differentiation with respect to time are also interchangeable, i.e,
n oT ∂  
δ d˙ = δ {d}T (5.13)
∂t
Using Equation 5.13, the first term of Equation 5.12 can be written as
ˆt2 n oT n o ˆt2 n o
˙ ˙ ∂
δ d [m] d dt = δ {d}T [m] d˙ dt
∂t
t1 t1

This term can be evaluated by using integration by parts as follows


ˆt2 n o n o t2 ˆt2 n o

δ {d} [m] d˙ dt = δ {d} [m] d˙
T T
− δ {d}T [m] d¨ dt (5.14)
∂t t1
t1 t1

At times t1 and t2 , the displacement must satisfy initial and final conditions, and therefore
no variation is allowed. This means that δ {d}T is 0 at these times. Therefore the first term
on right hand side of Equation 5.14 is 0. Substituting the remaining term in Equation
5.12 we obtain
ˆt2 n o ˆt2 ˆt2
− δ {d} [m] d¨ dt − δ {d} [k] {d} dt + δ {d}T {f } dt = 0
T T

t1 t1 t1
ˆt2  n o 
T ¨ T T
−δ {d} [m] d − δ {d} [k] {d} + δ {d} {f } = 0
t1

For this integral to be 0 for an 


arbitrary
n integrand,
o the integrand
 itself must vanish, i.e.,
δ {d} − [m] d¨ − [k] {d} + {f } = 0
T

For any arbitrary variation in displacement, the only insurance for this equation to hold
is
n o
− [m] d¨ − [k] {d} + {f } = 0
n o
⇒ [m] d¨ + [k] {d} = {f } (5.15)

This is the finite element equation for the element representing the dynamic equilibrium
condition. The equation is similar to that for a static system, with the addition
n o of one
extra force term in the left hand side of the equation. This force term [m] d¨ is known
as the inertia force of the element.

5.6 Coordinate transformation and system equations


The coordinate transformation equations developed in Section 2.10 apply also to dynamic
quantities such as mass matrix, velocity vector, and acceleration vector. If the element
coordinate system and global coordinate system are not the same, then a proper trans-
formation is needed before element equations are assembled to form system equations.
CHAPTER 5. FE FORMULATION OF DYNAMIC SYSTEMS 62

 0
If m and [m] represent mass matrices of an element in element and global coordinate
systems, respectively, and [T ] is the transformation matrix, the transformation equation
can be written as

[m] = [T ]T [m] [T ] (5.16)


When all the element equations are transformed to a common global coordinate system,
the equations can be assembled to from the finite element equation of the whole system.
The assembly procedure is identical to the one discussed in Section 2.9. Mass matrix
of the system is assembled from the mass matrices of all the elements, in the same way
that stiffness matrix is assembled. Once the assembly procedure is completed, the system
equation can be written as
n o
[M ] D̈ + [K] {D} = {F } (5.17)

5.7 Mass matrix of some common structural members


Mass matrix of an element can be created according to the developments in Section 5.3.
For an element, the mass matrix is obtained by evaluating the following integral
ˆ
[m] = ρ [N ]T [N ] dV (5.18)
Ve

Mass matrix obtained from Equation 5.18 is called as consistent mass matrix because it
is obtained by using the same shape function matrix as that used in creating the stiffness
matrix. In some practical applications, a simpler form of mass matrix is often used. The
simplest form is obtained by direct lumping of masses at the nodal degrees of freedom,
ignoring any cross coupling between different degrees of freedom. Such mass matrices are
called as lumped mass matrices, and are diagonal in structure. Consider a simple example
of a truss element of length L, cross sectional area A, and mass density ρ. The total mass
of the element is ρAL. A lumped mass matrix is obtained by assigning equal mass of
ρAL/2 to the two degrees of freedom. The diagonal terms are taken as zero. Thus the
lumped mass matrix of a truss element is
 
ρAL 1 0
[m] =
2 0 1
Use of lumped mass matrices are motivated by that fact that a diagonal mass matrix may
offer computational and storage advantages in certain situations. In the rest of this book,
only consistent mass matrices are considered.

5.7.1 Truss element


The mass matrix for a truss element is obtained by using the shape function matrix
derived in Section 3.3.
CHAPTER 5. FE FORMULATION OF DYNAMIC SYSTEMS 63

ˆ
[m] = ρ [N ]T [N ] dV
Ve
ˆ  
N1 
= ρA N1 N2 dx
N2
L
ˆ  
N1 N1 N1 N2
= ρA dx (5.19)
N1 N2 N2 N2
L

The integrals are evaluated as

ˆ ˆ 
x 2
N1 N1 dx = 1− dx
L
L L
ˆ  
2x x2
= 1− + 2 dx
L L
L
 L
x2 x3
= x− + 2
L 3L 0
L
=
3

ˆ ˆ 
x x
N1 N2 dx = 1− dx
L L
L L
ˆ  
x x2
= − dx
L L2
L
 2 L
x x3
= −
2L 3L2 0
L
=
6
ˆ ˆ  3 L
x2 x L
N2 N2 dx = 2
dx = 2
=
L 3L 0 3
L L

Replacing these integrals in Equation 5.19 we obtain


 
ρAL 2 1
[m] = (5.20)
6 1 2

5.7.2 Plane beam element


Consider next a plane beam element in the local xy plane (see Figure 4.4). The mass
matrix is obtained as
CHAPTER 5. FE FORMULATION OF DYNAMIC SYSTEMS 64

 
ˆ ˆ N1 N1 N1 N2 N1 N3 N1 N4
 N1 N2 N2 N2 N2 N3 N2 N4 
[m] = ρ [N ]T [N ] dV = ρA  N1 N3
 dx
N2 N3 N3 N3 N3 N4 
Ve L N1 N4 N2 N4 N3 N4 N4 N4
The integrals can be evaluated one at a time, obtaining the element of the matrix. An
example is shown below
ˆ ˆ  2
2x3 3x2
N1 N1 dx = 1+ 3 − 2 dx
L L
L L
ˆ  
4x6 9x4 4x3 12x5 6x2
= 1 + 6 + 4 + 3 − 5 − 2 dx
L L L L L
L
 
4 9 4 12 6
= L 1+ + + − −
7 5 4 6 3
 
60 + 189 − 210
= L
105
39L
=
105
The other terms can be evaluatedsimilarly, and the mass matrix 
is formed as
156 22L 54 −13L
ρAL  22L 4L 2
13L −3L2  
[m] = (5.21)
420  54 13L 156 −22L 
−13L −3L2 −22L 4L2

5.7.3 Plane frame element


For a plane frame element with 2 truss degrees of freedom, and 4 degrees of freedom
corresponding to a planar beam bending in the local xy plane, the mass matrix can be
created by assembling the mass matrices of truss and beam at the corresponding degrees
of freedom. The truss degrees of freedom are 1 and 4, while the beam degrees of freedom
are 2, 3, 5, and 6. Thus the mass matrix is obtained as
 1 
3
0 0 13 0 0
 0 13 11L
0 70 9 −13L 
 35 210 420 
 0 11L L2
0 13L −L2 
[m] = ρAL 
 1
210 105 420 140 
 6 0 0 1
6
0 0 
 0 9 13L
0 13 −11L 
70 420 35 210
−13L −L2 −11L L2
0 420 140
0 210 105

5.7.4 Space frame element


To be completed at a later date
Chapter 6

Free vibration of SDOF system

6.1 Introduction
The dynamic equation of motion represented by Equation 5.18 is a system of differential
equations. The solution of such equations can be obtained by using mathematical methods
of differential calculus. This field of mathematics is well developed, and standard solutions
for numerous forms of equations applicable to practical problems are available. Before
attempting to solve a system of differential equations, it is essential that one understands
the methods of solving a single differential equation. In this chapter we consider a special
case of Equation 5.19 where the system under study consists of a single degree of freedom
reducing the system of differential equation to a single differential equation. The physical
model corresponding to such a case is often called as a SDOF dynamic system. The study
of the dynamic response of a SDOF system is very useful because it provides us with
the foundation upon which the solution of Multi Degree oF Freedom (MDOF) systems
can be developed. Basic methods of solving a SDOF dynamic equations representing free
vibration problems are introduced. Energy dissipation mechanism in the form of vicious
damping is introduced next, along with the corresponding solution methods. Only free
vibration problems are considered in this chapter, and forced vibration analysis is covered
in the next chapter.

6.2 Free vibration of SDOF systems


6.2.1 System configuration
When the system under study has a single DOF, the matrix Equation 5.18 reduces to the
following form.

mü(t) + ku(t) = f (t) (6.1)


This equation the dynamic equilibrium of a system with mass m, and stiffness k. The
displacement of the system d and the force applied to it f are time dependent functions.
An idealized physical representation of such a system is shown in Figure 6.1. The elastic
property of the system is represented by a linearly elastic spring characterized by stiffness
(spring constant) of k . The inertial property of the system is represented by a rigid block
having mass m. The motion of the rigid mass is constrained by the frictionless rollers
in such a way that it can move only in simple translation, and a single displacement

65
CHAPTER 6. FREE VIBRATION OF SDOF SYSTEM 66

component completely defines its position at any given time The mechanical system, in
this setup, is therefore called as a Single Degree Of Freedom (SDOF) system. Any external
dynamic load, if applied to the system, is represented as a function of time f (t).

u(t)
k
m f(t)

Figure 6.1: A mechanical model of a SDOF dynamic system.

6.2.2 Free vibration equation and solution


If the external forces are not present, and the vibration is entirely caused by some distur-
bance (displacement and/or velocity) at the initial time, the equation of motion is of the
following form with specified initial displacement u0 and u̇0 and time t = 0.

mü(t) + ku(t) = 0, u(0) = u0 , u̇(0) = u̇0 (6.2)


Equation 6.2 is a second-order homogeneous differential equation with constant coeffi-
cients. The solution of this differential equation can be obtained by assuming the following
harmonic displacement function

u(t) = est (6.3)


where s is a constant to be determined. Substituting u = est and ü = s2 est in Equation
6.2, we obtain

ms2 + ks est = 0 (6.4)
Since est = 0 is trivial, the non-trivial solution results in

ms2 + k = 0 (6.5)
which is called the characteristic equation of motion with its roots given by
r
k
s1,2 = ±i (6.6)
m
q
where i is the imaginary number. The quantity m k
is replaced by ω, called as the natural
frequency of the system and thus s = ±iω. This indicates that es1 t and es2 t both satisfy
Equation 6.2. If this is the case, any linear combination of these functions must also be
a solution and can be written, after introducing two constants A1 ,A2 , as

u(t) = A1 eiωt + A2 e−iωt (6.7)


By using the Euler’s equations eix = cos x + i sin x , e−ix = cos x − i sin x , the exponential
functions can be converted to trigonometric functions.
CHAPTER 6. FREE VIBRATION OF SDOF SYSTEM 67

u(t) = A1 eiωt + A2 e−iωt


= A1 (cos ωt + i sin ωt) + A2 (cos ωt − i sin ωt)
= (A1 + A2 ) cos ωt + (A1 − A2 ) i sin ωt
= A cos ωt + B sin ωt (6.8)

The constants A and B are to be determined from the initial conditions. Using u(t) = u0
at t = 0 in Equation 6.8 results in

A = u0 (6.9)
The velocity of the system is obtained by differentiating Equation 6.8 with t.

u̇(t) = −Aω sin ωt + Bω cos ωt (6.10)


Using u̇ = u̇0 at t = 0 in Equation 6.10 we get
u̇0
B= (6.11)
ω
Combining Equations 6.8, 6.9, and 6.11 the solution of the displacement response can be
written as follows.
u˙0
u(t) = u0 cos ωt + sin ωt (6.12)
ω
This solution represents a simple harmonic motion (SHM) with an angular frequency of
ω. It is easily verified from Equation 6.12 that

u(t) = u(t + ) (6.13)
ω
This simply means that the motion repeats itself every 2π
ω
second, and therefore the period
of vibration, commonly known as natural period of vibration, is given by

T = (6.14)
ω
This quantity represents the time taken for the SDOF system to complete one cycle of its
vibration. The inverse of this quantity gives a measure of how many cycles are completed
in one second, and is called as the natural frequency of the system
1
f= (6.15)
T
The frequency is measure in units of hertz (Hz), where one Hz represents one cycle per
second. The quantities f and ω are often referred to as undamped natural frequency of
the SDOF; the distinction whether it is the angular frequency or the cyclic frequency is
merely made by the specified units of measurement, which is rad/s for ω. The motion
described by Equation 6.12 is plotted in Figure 6.2 for different combinations of initial
displacements and velocities. The natural period is taken as 1s.
CHAPTER 6. FREE VIBRATION OF SDOF SYSTEM 68

1.5

1
Displacement (m)

0.5

u0 = 1, u̇ 0 = 0
u0 = 1, u̇ 0 = 5
0
u0 = 1, u̇ 0 = −5
u0 = 0, u̇ 0 = 5

−0.5

−1

−1.5
0 0.5 1 1.5 2 2.5 3
Time (s)

Figure 6.2: Free vibration response of a SDOF system for different initial conditions.

6.3 Damped free vibration of SDOF systems


The vibration of the SDOF system shown in Figure 6.2 continues for ever with the same
amplitude (peak displacement) and frequency. This is, however, not a realistic scenario in
the sense that real-life mechanical systems dissipate energy which results in the decrease in
amplitude of motion. Energy dissipation causes the motion to cease after a finite time, and
such a mechanism is called as damping. Damping is a very complex phenomenon, and an
exact physical model to quantify damping mechanism does not exist. However, based on
experience and experiments, empirical models of damping have been established for many
structural and mechanical systems. The most commonly used model of damping is the
so-called viscous damping model. According to this model, a moving system experiences
a resisting force, called as damping force, which is linearly proportional to the velocity of
motion. The proportionality constant between the damping force and the velocity of the
system is called as damping coefficient.

6.3.1 System configuration


To model the damping characteristics of the SDOF system shown in Figure 6.1, a viscous
dashpot is attached to the rigid block. The configuration is schematically shown in Figure
6.3. The dashpot exerts a force to the moving mass; the magnitude of the force is
proportional to the velocity of the moving mass, and it acts in a direction opposite to the
direction of motion. The damping coefficient is c.
CHAPTER 6. FREE VIBRATION OF SDOF SYSTEM 69

u(t)
k
m f(t)
c

Figure 6.3: A mechanical model of a SDOF dynamic system with viscous damping.

6.3.2 Damped free vibration equation and solution


The equation of motion is similar to Equation 6.2 with damping force added to the left
side of the equation.

mü(t) + cu̇(t) + ku(t) = 0, u(0) = u0 , u̇(0) = u̇0 (6.16)


Dividing Equation 6.16 by m and introducing ω 2 = k/m we obtain.
c
ü(t) + u̇(t) + ω 2 u(t) = 0 (6.17)
m
The solution to this homogeneous second-order differential equation with constant coef-
ficients can be assumed to be of the exponential form u(t) = est . Substituting for u(t),
u̇(t) = sest , and ü(t) = s2 est in Equation 6.17 we obtain
 c 
s2 + s + ω 2 est = 0 (6.18)
m
from which the non-trivial solution is obtained as
c
s2 + s + ω2 = 0 (6.19)
m
which is called as the characteristic equation of motion of the damped SDOF system.
This is a quadratic equation in s and the roots are given by
r
c c 2
s1,2 = − ± − ω2 (6.20)
2m 2m
Depending on the sign of the discriminant (the quantity inside the radical) the roots of
the characteristic equation can be (i) real and equal (ii) real and distinct, or (iii) complex
conjugates of each other. Each of these cases results in in different solutions for the
displacement of the SDOF as a function of time, and they are considered separately in
the following.

6.3.2.1 Case 1 : Critically damped system


If the discriminant is equal to zero, the two roots obtained from Equation 6.20 are real
and equal. The damping coefficient that corresponds to this solution is called as critical
damping coefficient, denoted as cc , and the corresponding motion is called as critically
damped motion. Equating the discriminant to zero and replacing c by cc we obtain
CHAPTER 6. FREE VIBRATION OF SDOF SYSTEM 70

 c 2
c
− ω2 = 0
2m
⇒ cc = 2mω (6.21)

Equation 6.21 indicates that the critical damping coefficient depends entirely on the mass
and the angular frequency of the SDOF system. The roots of the characteristic equation
are thus obtained from Equation 6.20 as
cc
s = s1 = s2 = − = −ω (6.22)
2m
Then the two independent solutions of Equation 6.16 are u(t) = est , and u(t) = test .
Therefore, a linear combination of these two independent solutions must also be a solution
of the differential equation, and the solution can be written as

u(t) = A1 e−ωt + tA2 e−ωt (6.23)


The constants A1 and A2 are determined from the initial conditions of motion. Using
u(0) = u0 and t = 0 in Equation 6.23 results inA1 = u0 . Differentiating Equation 6.23
with respect to t results in the following expression for velocity.

u̇(t) = −ωA1 e−ωt − ωtA2 e−ωt + A2 e−ωt (6.24)


Using u̇(0) = u̇0 and t = 0 in Equation 6.24 results in the following expression for A2 .

A2 = u0 ω + u̇0 (6.25)
The final solution is obtained by substituting the values of A1 and A2 in Equation 6.23
resulting in

u(t) = (u0 + (ωu0 + u̇0 ) t) e−ωt (6.26)


The solution described by Equation 6.26 is plotted in Figure 6.4 for different initial con-
ditions, and ω = 2π.
CHAPTER 6. FREE VIBRATION OF SDOF SYSTEM 71

1.2
u0 = 1, u̇0 =0
u0 = 1, u̇0 =5
1 u0 = 1, u̇0 = −10
u0 = 0, u̇0 =5
0.8
Displacement (m)

0.6

0.4

0.2

−0.2
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Time (s)

Figure 6.4: Critically damped free vibration response of a SDOF system for different
initial conditions.

The motion depicted in Figure 6.4 is not oscillatory (not vibrating). Starting from zero
time, the displacement builds up to a maximum or a minimum value that depends on
the initial displacement and initial velocity of the system. After that, the motion rapidly
decays. The rate of decay is exponential in time as indicated by Equation 6.26. Because
the amount of damping is just sufficient to inhibit vibration of the system, the corre-
sponding damping coefficient is called as critical damping coefficient. The rate of decay
of motion depends on the angular frequency of the system. In order to illustrate this
effect, the responses of two SDOF systems with the same initial conditions, but different
angular frequencies are compared in Figure 6.5 below. The SDOF system with smaller
angular frequency displays larger peak motion than the one with larger angular frequency.
The decay of motion from its peak value is faster for systems which have larger angular
frequency. This is an indication that high frequency motion of a critically damped system
decay faster than low frequency motion.
CHAPTER 6. FREE VIBRATION OF SDOF SYSTEM 72

3
ω = 1 rad/s
ω = 2 rad/s
2.5
Displacement (m)

1.5

0.5

0
0 2 4 6 8 10 12 14
Time (s)

Figure 6.5: Critically damped motion of two SDOF systems with identical initial condi-
tions, but different angular frequencies.

6.3.2.2 Case 2 : Overdamped system


When the damping coefficient is greater than the critical damping coefficient, the motion
of the system is said to be overdamped. The ratio between the actual damping coefficient
and the critical damping coefficient is called as damping ratio, and is denoted by ζ.
c c
ζ= = (6.27)
cc 2mω
The roots of the characteristic equation of motion (Equation 6.20) for overdamped systems
are real and distinct, and they can be obtained by replacing c/2m = ζω.

p
s1,2 = −ζω ± ζ 2ω2 − ω2
p
= −ζω ± ω ζ 2 − 1
= −ζω ± ωD (6.28)
p
where ωD ≡ ω ζ 2 − 1 is called as the damped angular frequency of the system. Since the
roots of the equation are distinct, the independent solutions of the differential equation of
motion are u(t) = es1 t and u(t) = es2 t . Introducing two constants A1 and A2 the solution
can be expressed as a linear combination of these two independent solution.

u(t) = A1 e(−ζω+ωD )t + A2 e(−ζω−ωD )t



= e−ζωt A1 eωD t + A2 e−ωD t (6.29)

The exponential functions in the above equation can be expressed as hyperbolic trigono-
metric functions by using the relations ex = cosh x + sinh x , and e−x = cosh x − sinh x.

u(t) = e−ζωt [(A1 + A2 ) cosh ωD t + (A1 − A2 ) sinh ωD t] (6.30)


CHAPTER 6. FREE VIBRATION OF SDOF SYSTEM 73

The arbitrary constants A1 + A2 and A1 − A2 are denoted as constants A and B, and the
solution written as.

u(t) = e−ζωt [A cosh ωD t + B sinh ωD t] (6.31)


Using the initial condition u(0) = u0 at t = 0 in Equation 6.31 results in A = u0 .
Differentiating Equation 6.31 with t, the velocity of motion can be expressed as

u̇(t) = e−ζωt [AωD sinh ωD t + BωD cosh ωD t] − ζωe−ζωt [A cosh ωD t + B sinh ωD t] (6.32)

Using the initial condition u̇(0) = u̇0 at t = 0 in Equation 6.32 we get

BωD − Aζω = u̇0


u̇0 + u0 ζω
⇒B = (6.33)
ωD
Replacing for the constants A and B in Equation 6.31 gives the final solution.
   
u̇0 + u0 ζω
u(t) = e−ζωt
u0 cosh ωD t + sinh ωD (6.34)
ωD
The solution is plotted in Figure 6.6for different initial conditions.

1.4
u0 = 1, u̇0 =0
u0 = 1, u̇0 =5
1.2 u0 = 1, u̇0 = −10
u0 = 0, u̇0 =5
1
Displacement (m)

0.8

0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Time (s)

Figure 6.6: Motion of an overdamped system; different initial conditions as depicted on


the legend are plotted, the damping ratio is ζ = 1.1 , and the frequency of the system is
ω=2π rad/s.

The overdamped motion, like the critically damped motion, is non-oscillatory. The dis-
placement builds up to maximum amplitude from where it decays rapidly and comes to
zero after a certain time. The rate of decay depends on both the angular frequency of
the system and the damping ratio. A comparison of the motion of SDOF systems with
identical angular frequencies and initial conditions, but different damping ratios is shown
in Figure 7.
CHAPTER 6. FREE VIBRATION OF SDOF SYSTEM 74

1.4
ζ = 1.00
ζ = 1.20
1.2 ζ = 1.50
ζ = 2.00
1
Displacement (m)

0.8

0.6

0.4

0.2

0
0 0.5 1 1.5 2 2.5 3
Time (s)

Figure 6.7: Motion of critically damped and overdamped systems; the initial conditions
are the same for all the systems, the period of all the systems is 1s, and different damping
ratios are considered as shown in the legend; note that the green line corresponds to the
critically damped system whose solution is obtained from Equation 6.26.

Figure 6.7 shows that the peak amplitude decreases as damping ratio is increased above
the critical value. The rate of decay, on the other hand, decreases as damping ratio
is increased above the critical level. It should also be noted that the rate of decay of
the critically damped motion is slower than that of overdamped systems. Thus for two
systems starting from the same initial conditions, overdamped motion achieves a smaller
peak and takes longer time to come to a state of rest than the critically damped system.

6.3.2.3 Case 3: Underdamped system


When the damping coefficient is less than the critical value, the motion is said to be
underdamped. Underdamped systems are characterized by a damping ratio less than
one, i.e. ζ < 1. The roots of the characteristic equation of motion (Equation 6.20) for
underdamped systems are complex and distinct (they are complex conjugates), and they
can be directly obtained from by replacing c/2m = ζω as
p p
s1,2 = −ζω ± ζ 2 ω 2 − ω 2 = −ζω ± iω 1 − ζ 2 = −ζω ± iωD (6.35)
p
where ωD ≡ ω 1 − ζ 2 is called as the damped angular frequency of the underdamped
system. Note that the damped frequencies of overdamped and underdamped systems
are defined in slightly different ways (compare equations 6.28 and 6.35). Since the two
roots are distinct, the independent solutions of the differential equation of motion are
u(t) = es1 t , and u(t) = es2 t . Any linear combination of the two solutions is also a solution
of the differential equation, which gives us

u(t) = A1 es1 t + A2 es2 t


= A1 e(−ζω+iωD )t + A2 e(−ζω−iωD t)

= e−ζωt A1 eiωD t + A2 e−iωD t (6.36)
CHAPTER 6. FREE VIBRATION OF SDOF SYSTEM 75

The complex exponential in the above equation can be converted to trigonometric func-
tions by suing the Euler’s equations.

u(t) = e−ζωt [(A1 + A2 ) cos ωD t + (A1 − A2 ) sin ωD t] (6.37)


Introducing another set of constants A = A1 + A2 and B = A1 − A2 , we get.

u(t) = e−ζωt (A cos ωD t + B sin ωD t) (6.38)


Using the initial condition u(0) = u0 at t = 0 in Equation 6.38 results in A = u0 .
Differentiating Equation 6.38 with t results in the following equation for velocity of motion.

u̇(t) = e−ζωt (−AωD sin ωD t + BωD cos ωD t) − ζωe−ζωt (A cos ωD t + B sin ωD t) (6.39)

Using the initial condition u̇(0) = u̇0 and t = 0 in Equation 6.39 we obtain the following
relation.

BωD − Aζω = u̇0


u̇0 + u0 ζω
⇒B = (6.40)
ωD
Using the constants A and B in Equation 6.38 gives the final solution of motion of un-
derdamped systems.
   
u̇0 + u0 ζω
u(t) = e −ζωt
u0 cos ωD t + sin ωD t (6.41)
ωD
The motion described by the above equation is oscillatory. The expression inside the
parenthesis represents a simple harmonic motion with frequency ωD . The exponential
function e−ζωt causes the amplitude of the harmonic motion to decay with time, eventually
bringing the system to rest. The amplitude of the harmonic part of the motion is given
by
s  2
u̇0 + u0 ζω
ρ = (u0 ) +2
(6.42)
ωD
The displacement solution given by Equation 6.41 is plotted in Figure 8. The displacement
oscillates with an angular frequency of ωD , and the corresponding period of vibration is
TD = 2π/ωD . The red and the green lines represent the envelope of vibration which
decays exponentially.
CHAPTER 6. FREE VIBRATION OF SDOF SYSTEM 76

1.5
u(t)
ρe−ζωt
1 −ρe−ζωt
Displacement (m)

0.5

−0.5

−1

−1.5
0 1 2 3 4 5 6 7 8
Time (s)

Figure 6.8: Vibration of an underdamped system (ζ = 0.05), undamped period of vibra-


tion T = 1s, initial displacement 1m, and initial velocity 5 m/s; the red and the green
line indicate the decay of the amplitude of motion, the period of vibration is equal to the
damped period TD = √2π 2 .
ω 1−ζ

6.3.3 Effect of damping


Viscous damping leads to the decay of free vibration. It also leads to positive effects
(reduction in amplitude) when added to systems with external excitation (forces). For
these reasons, viscous damping is often artificially added to many mechanical systems.
Many systems are designed with less than critical viscous damping because of its favorable
effect in reducing amplitudes of forced vibrations. When a system is subject to free
vibration only, it may be designed with critical damping or over critical damping to
inhibit oscillatory motion. A critically damped system returns to rest quicker than an
overdamped system. Thus recoil mechanisms in guns are designed with critical damping
to allow rapid firing. Automobile suspensions are often subject to both free and forced
vibrations. Free vibrations occur when the vehicle is subject to sudden change in road
contour. In this case, the shock absorbers should have a damping ratio close to 1 (critically
damped). However, if the vehicle is traveling on a bumpy road, the vehicle is subject to
a possible random external excitation. In this case the system should be underdamped.
For these reasons vehicle shock absorbers are often designed to be self-adaptive. The
damping ratio of structural systems is much less than one. Common reinforced concrete
buildings can have a damping ratio in the order of 0.02 to 0.05. It is very uncommon for
a structural system to have damping ratios greater than 0.2, unless additional artificial
damping mechanisms are installed intentionally. Such supplemental damping is gaining
popularity to reduce the vibrations of structural systems caused by dynamic excitations
such as wind vibrations, or earthquake-induced vibrations. From structural engineering
perspective, underdamped motion is therefore of the greatest interest. In underdamped
conditions, presence of larger damping causes the system to come back to equilibrium
position more rapidly. This effect is shown in Figure 9 for two SDOF systems with
different damping ratios, but identical otherwise.
CHAPTER 6. FREE VIBRATION OF SDOF SYSTEM 77

1.5
ζ = 0.02
ζ = 0.10
1
Displacement (m)

0.5

−0.5

−1

−1.5
0 1 2 3 4 5 6 7 8
Time (s)

Figure 6.9: Effect of damping on the underdamped vibration of SDOF systems; both
systems start with the same initial displacement of 1m and initial velocity of 5m/s; the
undamped natural period of both the systems is 1s; their damping ratios are 2% and 10%
of critical; note that the effect of damping in free vibration is not so much in reducing the
maximum displacement but in increasing the rate at which the oscillations decay bringing
the system back to rest.

6.3.4 Measurement of damping


The true damping characteristics of typical structural systems are very complex and
difficult to define. However, it is common practice to express the damping of such real
systems in terms of equivalent viscous-damping ratios which show similar decay rates
under free vibration conditions. It is therefore possible to experimentally evaluate the
damping ratio by recording the free vibration of structures. For a SDOF system, a
framework used to evaluate the viscous damping ratio is explained next. Consider the
vibration of a structural system due to a given initial displacement and zero initial velocity.
The displacement response is then obtained from Equation 6.41 by setting u̇0 = 0 .
   
u0 ζω
u(t) = e−ζωt
u0 cos ωD t + sin ωD t (6.43)
ωD
The displacement is maximum at t = 0 from where it oscillates with exponentially de-
caying amplitudes. The first positive peak occurs at t = 0, and the second peak occurs
after one complete cycle of vibration which is equal to the damped period of the system,
TD = 2π/ωD . This shows that any consecutive peaks n and (n+1) are separated by a
time of 2π/ωD . An example of such a motion is shown in Figure 6.10, where the second
and the sixth peaks are marked.
CHAPTER 6. FREE VIBRATION OF SDOF SYSTEM 78

0.8

0.6
Displacement (cm)

0.4

0.2

−0.2

−0.4

−0.6

−0.8

−1
0 1 2 3 4 5 6 7 8 9 10
Time (s)

Figure 6.10: An example of free vibration of a SDOF system with initial displacement
1m, the second and the sixth peaks are marked with red dots.

If tn is the time of nth peak, the time of (n+1)th peak is tn +2π/ωD . Let the corresponding
displacements at these times are equal to and un and un+1 , respectively. From Equation
6.43 we get
   
u0 ζω
un = e −ζωtn
u0 cos ωD tn + sin ωD tn (6.44)
ωD

   
u0 ζω
 
−ζω tn + ω2π
un+1 = e D u0 cos (ωD tn + 2π) + sin (ωD tn + 2π) (6.45)
ωD

Dividing Equation 6.44 by Equation 6.45 gives us


un
 
2πζω
−ζω − ω2π
=e D =e ωD
(6.46)
un+1
Taking
p the natural logarithm of both sides of this equation and substituting ωD =
ω 1 − ζ 2 , we obtain the so-called logarithmic amplitude ratio, δ, defined as
 
un 2πζω 2πζ
δ = ln = p =p (6.47)
un+1 ω 1 − ζ2 1 − ζ2
The solution of this equation gives the required value of damping ratio ζ. Since this
equation is nonlinear in ζ an analytical solution is not available,
p and numerical procedure
needs to be adopted. If the damping ratio is small, then 1 − ζ ≈ 1, and damping ratio
2

can be obtained as
δ
ζ≈ (6.48)

In practical applications, it might be more convenient to consider peaks which are well
separated rather than two consecutive peaks. If nth and (n+m)th peaks are considered,
CHAPTER 6. FREE VIBRATION OF SDOF SYSTEM 79

and the corresponding displacements are un and un+m , respectively, damping ratio can
be computed as
 
un
ln un+m
ζ= (6.49)
2πm
If we consider the example shown in Figure 6.10 above, (the second peak), and (the sixth
peak is 4 cycles apart from the second peak, i.e., m = 4). The corresponding displacements
at these peaks are 0.73 m and 0.21 m, respectively. From Equation 6.49we compute the
damping ratio as
 

ln uu26 ln 0.73
0.21
ζ= = = 0.0496
2π(4) 2π(4)
The value of ζ used in generating Figure 6.10 is 0.005, which is almost equal to the
computedpvalue, the slight discrepancy is due to the approximation involved in neglecting
the term 1 − ζ 2 , which is satisfactory for small damping ratios.
Chapter 7

Forced vibration of SDOF systems

7.1 Introduction
The previous chapter dealt with SDOF systems vibrating due to some disturbance at
the initial time, without any sustained external excitation. In most practical situations,
engineers encounter vibration problems due to some external forces that vary with time.
This chapter develops the solution of dynamic equation of motion including external
forces. Solutions to vibration problems when the applied loads are harmonic in nature
is presented first. Solution methods that are applicable to an arbitrarily varying force is
introduced in the form of unit impulse response function, and Duhamel’s integral. Finally,
numerical methods for solving the forced vibration problems is introduced.

7.2 Response to harmonic excitations


Response of SDOF systems to harmonic excitations is of great interest in structural dy-
namics due to two important reasons. Firstly, such excitations are encountered in engi-
neering analysis, for example, forces due to unbalanced rotating machinery. Secondly, an
understanding of the dynamic response to harmonic excitations is useful in formulating
the response to other types of excitations. Since any periodic force can be decomposed
into harmonic functions of time by using the Fourier analysis, the harmonic solution along
with the principle of superposition can be used to find the response of a system to periodic
forces, at least for linearly elastic systems.
A harmonic force can be represented as f (t) = f0 sin ωt,
¯ or f (t) = f0 cos ω̄t where f0
is the amplitude of the force and its frequency ω̄ is called as the excitation frequency or
forcing frequency. The response of SDOF systems to sinusoidal forces will be derived in
detail in this section for both undamped as well as damped systems. The response to
cosine forces can be derived similarly, and only the final results are presented.

7.2.1 Harmonic response of undamped systems


An undamped SDOF system is subjected to a harmonic load f (t) = f0 sin ωt ¯ . The
equation of motion of the system is obtained by replacing the force term in Equation 6.1
by the applied sinusoidal force.

mü(t) + ku(t) = f0 sin ω̄t (7.1)

80
CHAPTER 7. FORCED VIBRATION OF SDOF SYSTEMS 81

This equation is to be solved for given initial displacement and velocity, u0 and u̇0 respec-
tively, at time t = 0. Since this is a non-homogeneous differential equation, the complete
solution is obtained by adding the complementary solution (solution of the homogenous
equation, or free vibration problem with the right hand side set to 0), and the particular
solution. The particular solution of this linear second-order differential equation is of the
form

up (t) = C sin ω̄t (7.2)


where C is a constant to be subsequently determined. Differentiating Equation 7.2 twice
with respect to time gives an expression for acceleration as üp (t) = −C ω̄ 2 sin ω̄t . Substi-
tuting up (t) and üp (t) in Equation 7.1 and using ω 2 = k/m , results in
f0 1
C=  (7.3)
k 1 − ω̄ 2
ω
The ratio of excitation frequency to the natural frequency of the system is called as
frequency ratio and denoted as β = ω̄/ω. Using this notation, the particular solution can
be written down as
 
f0 1
up (t) = sin ω̄t (7.4)
k 1 − β2
The complementary solution of Equation 7.1 is the solution of the homogeneous equation
which is exactly the solution of undamped free vibration, and is available as (see Equation
6.8)

uc (t) = A cos ωt + B sin ωt (7.5)


where A and B are constants to be determined from the initial conditions. The complete
solution is then obtained by adding the complimentary solution to the particular solution.
 
f0 1
u(t) = A cos ωt + B sin ωt + sin ω̄t (7.6)
k 1 − β2
The velocity of motion is obtained by differentiating the displacement with respect to
time.
 
f0 1
u(t) = −Aω sin ωt + Bω cos ωt + ω̄ cos ω̄t (7.7)
k 1 − β2
Using t = 0 and u(t) = u0 in Equation 7.6 results in A = u0 . Using t = 0 and u̇(t) = u̇0
in Equation 7.7 gives the following result.
 
u̇0 f0 β
B= − (7.8)
ω k 1 − β2
Thus the constants are determined, and Equation 7.6 can be used to determine the dis-
placement of the system. The solution given by Equation 7.8 is plotted in Figure 7.1.
The solution contains two distinct vibration components (1) the particular solution which
oscillates at the forcing frequency; and (2) the complimentary solution which oscillates
at the natural frequency of the system. The first of these is called as the steady-state
vibration. The latter is the transient vibration, and it depends entirely on the initial
conditions. In Figure 1 the dashed blue line represents the transient vibration component
CHAPTER 7. FORCED VIBRATION OF SDOF SYSTEMS 82

while the dashed red line corresponds to the steady state vibration component. The total
solution is the sum of these two components. The transient component for undamped
systems seems to continue forever. In real systems damping makes the transient vibration
to decay with time, and after a finite time, only steady-state vibration is present.

3
uc
up
uc + up
Displacement (m) 2

−1

−2
0 2 4 6 8 10
Time (s)

Figure 7.1: Harmonic response of an undamped SDOF system, natural period of the
SDOF is 1s, frequency ratio is 0.2, f0 /k = 1m, u0 = 1m, and u̇0 = 5 m/s.

7.2.1.1 Steady state response and displacement response factor


Because transient response of realistic systems is eventually damped out, the steady-state
response is of more interest to engineering problems. Neglecting the transient response
from Equation 7.6 we get the steady-state response as
 
f0 1
u(t) = sin ω̄t (7.9)
k 1 − β2
If inertial effects are not considered, the static solution is obtained from ust (t)k = f0 sin ωt
¯
. The ratio between the dynamic and the static displacement is called as the displacement
response factor and is given by

u(t) 1
Rd ≡ = (7.10)
ust (t) 1 − β2
This ratio indicates the amplification of motion due to dynamic effects, and is depen-
dent only on the frequency ratio β. The variation of displacement response factor with
frequency ratio is shown in Figure 7.2. Following observations can be made from the
figure.

1. β  1
In this case the loading frequency is smaller than the natural frequency of the system.
In the limit that the loading frequency is zero (the loading period is infinite, which
means static load), the displacement response factor is equal to 1, which means that
the response is static. In such situations, static analysis is sufficient to compute the
response of the system. Also note that the displacement response factor is positive
which means that the force and displacement are in phase (increasing force causes
increasing displacement and vice versa).
CHAPTER 7. FORCED VIBRATION OF SDOF SYSTEMS 83

2. β  1
In this case, the loading
√ frequency is much larger than the natural frequency of the
system. When β > 2, |Rd | < 1 which indicates that the dynamic response is
smaller than the static response. In the limit that the loading frequency is infinitely
larger than the natural frequency of the system, the response of the system is 0.
These results indicate that a system does not respond significantly to a load whose
frequency is much larger than its natural frequency. Also note that the displacement
response factor is positive which means that the force and displacement are out of
phase (increasing force causes decreasing displacement and vice versa).

3. β ≈ 1
When the loading frequency is very close to the natural frequency of the system, the
displacement response factor is very large. In the limit that the two frequencies are
equal, the dynamic response very large, and the phenomenon is called as resonance.
In these situations, the solution derived above is no longer valid, because when
ω ≈ ω̄, the solution goes to infinity, and a limiting case needs to be derived.

1
Rd

−1

−2

−3

−4

−5
0 0.5 1 1.5 2 2.5 3
ω̄
β= ω

Figure 7.2: Displacement response factor as function of frequency ratio.

7.2.1.2 Resonance
When ω = ω̄ , the solution given by Equation 7.6 needs to be evaluated at the limit
ω → ω̄ (Using L’Hopital’s rule).
CHAPTER 7. FORCED VIBRATION OF SDOF SYSTEMS 84

   
u̇0 f0 β f0
u(t) = lim u0 cos ωt + − sin ωt + sin ω̄t
ω→ω̄ ω k (1 − β 2 ) k (1 − β 2 )
   
u̇0 f0 ω̄ω f0 ω 2
= lim u0 cos ωt + − sin ωt + sin ω̄t
ω→ω̄ ω k (ω 2 − ω̄ 2 ) k (ω 2 − ω̄ 2 )
   
u˙0 f0 ω ω sin ω̄t − ω̄ sin ωt
= u0 cos ωt + sin ωt + lim
ω k ω + ω̄ ω→ω̄ ω − ω̄
   
u̇0 f0 ω sin ω̄t − ω̄t cos ωt
= u0 cos ωt + sin ωt + lim
ω k ω + ω̄ ω→ω̄ 1
u̇0 f0
= u0 cos ωt + sin ωt + (sin ωt − ωt cos ωt) (7.11)
ω 2k
If the system is initially at rest the complementary solution vanishes from Equation 7.11,
in which case we have the following result.
f0
(sin ωt − ωt cos ωt)
u(t) = (7.12)
2k
The solution given by Equation 7.12 is plotted in Figure 3. Here, the response of the
system grows constantly, and it becomes infinite after an infinitely long time. The rate
at which the displacement amplitude grows is given by the envelope shown with red lines
in Figure 3. The slope of these lines can be easily computed from Equation 7.12 and is
found to be πfk 0 .

40

30

20
Displacement (m)

10

−10

−20

−30

−40
0 2 4 6 8 10
Time (s)

Figure 7.3: Resonant response of an undamped SDOF system: T = 1s and f0 /k = 1m.

7.2.2 Harmonic response of damped systems


Consider a SDOF system with viscous damping coefficient c and subjected to a harmonic
excitation f (t) = f0 sin ω̄t. The differential equation of motion of the system is
CHAPTER 7. FORCED VIBRATION OF SDOF SYSTEMS 85

mü(t) + cu̇(t) + ku(t) = f0 sin ω̄t (7.13)


Dividing by m , and using ω 2 = k/m, c/m = 2ζω we get
f0
ü(t) + 2ζω u̇(t) + ω 2 u(t) =
sin ω̄t (7.14)
m
The particular solution of this differential equation is of the form

up (t) = C cos ω̄t + D sin ω̄t (7.15)


from which the velocity and displacement can be obtained by successive differentiation to
obtain the following.

u̇p (t) = −ω̄C sin ω̄t + ω̄D cos ω̄t (7.16)

üp (t) = −ω̄ 2 C cos ω̄t − ω̄ 2 D sin ω̄t (7.17)


Substituting Equations 7.15, 7.16, and 7.17 in Equation 7.14 we get

  f0
cos ω̄t −ω̄ 2 C + 2ζω ω̄D + ω 2 C + sin ω̄t −ω̄ 2 D − 2ζω ω̄C + ω 2 D = sin ω̄t (7.18)
m
Equating the coefficients of sin ω̄t and cos ω̄t on the two sides of Equation 7.18 results in
the following two equations.

− C ω̄ 2 + 2ζω ω̄D + ω 2 C = 0
⇒ −Cβ 2 + 2ζβD + C = 0
⇒ C(1 − β 2 ) + D(2ζβ) = 0 (7.19)

f0
− ω̄ 2 D − 2ζω ω̄C + ω 2 D =
m
f0
⇒ −Dβ 2 − 2ζβC + D =
mω 2
f0
⇒ D(1 − β 2 ) + C(−2ζβ) = (7.20)
k
Solving Equations 7.19 and 7.20 simultaneously we get

   
2ζβ 1 − β2 f0
D −D =−
1 − β2 2ζβ 2ζβk
( )
2
(1 − β 2 ) − (2ζβ)2 f0
⇒D =
1 − β2 k
 
f0 1 − β2
⇒D= (7.21)
k (1 − β 2 )2 + (2ζβ)2
CHAPTER 7. FORCED VIBRATION OF SDOF SYSTEMS 86

D (2ζβ)
C = −
1 − β2
 
f0 −2ζβ
⇒C = (7.22)
k (1 − β 2 )2 + (2ζβ)2

The complimentary solution is the solution of free vibration problem, which for undamped
system is obtained from Equation 6.38. The total solution is obtained by adding the
complimentary and the particular solutions.

u(t) = e−ζωt (A cos ωD t + B sin ωD t) + C cos ω̄t + D sin ω̄t (7.23)


The velocity response is obtained by differentiation with respect to time.

u̇(t) = e−ζωt (−AωD sin ωD t + BωD cos ωD t)


− ζωe−ζωt (A cos ωD t + B sin ωD t) − C ω̄ sin ω̄t + Dω̄ cos ω̄t (7.24)

Using the initial condition t = 0, and u(t) = u0 in Equation 7.23 gives the following.

A = u0 − C (7.25)
Using the initial condition t = 0, and u̇(t) = u̇0 in Equation 7.24 gives the following.

BωD − Aζω + Dω̄ = u̇0


u̇0 − Dω̄ + Aζω
⇒B = (7.26)
ωD
Thus a complete solution is obtained for underdamped harmonic response of a SDOF
system. This solution is plotted in Figure 7.5 for selected parameter values as indicated
in the figure caption. The total solution consists of two distinct components: steady-state
response and transient response. The transient response is the free vibration response
which decays exponentially with time. The steady-state response is the particular solution,
plotted as red line, and it oscillates at the loading frequency. After sufficient time, the
transient response comes to zero due to the exponential decay, and the total response is
just the steady-state response.
CHAPTER 7. FORCED VIBRATION OF SDOF SYSTEMS 87

3
uc
2.5 up
uc + up
2

Displacement (m)
1.5

0.5

−0.5

−1

−1.5

−2
0 2 4 6 8 10 12
Time (s)

Figure 7.4: Underdamped harmonic vibration of a SDOF system: T = 1s, β = 0.2,


ζ = 0.05, f0 /k = 1m ,u0 = 1m , and u̇0 = 0 are used for illustration; the blue, red, and
black lines represent the transient, steady-state, and total response, respectively.

7.2.2.1 Steady state response and displacement response factor


The steady state response is the particular solution of the damped harmonic response,
which from Equation
√ 7.23 is u(t) = C cos ω̄t+C sin ω̄t. The amplitude of the this dynamic
response is ρ = C 2 + D2 . If inertial effects are neglected, the static response is simply
u(t) = (f0 /k) sin ω̄t , and the amplitude of static response is f0 /k . The ratio between the
amplitudes of dynamic and static response is called as the displacement response factor
(or dynamic amplification factor).


C 2 + D2
Rd =
f0/k
s
1
= (7.27)
(1 − β 2 )2+ (2ζβ)2

The deformation response factors are shown for different damping ratios in Figure 7.5.
CHAPTER 7. FORCED VIBRATION OF SDOF SYSTEMS 88

5
ζ = 0.02
ζ = 0.10
4.5 ζ = 0.20
ζ = 0.70
4 ζ = 0.99

3.5

3
Rd

2.5

1.5

0.5

0
0 0.5 1 1.5 2 2.5 3
ω̄
β= ω

Figure 7.5: Deformation response factor as a function of frequency ratio and damping
ratio.

The effect of damping in the dynamic response depends on the frequency ratio, and is
summarized below

1. If the frequency ratio is much smaller than 1, β  1, Rd is only slightly larger than
1, and is essentially independent of the damping ratio. This implies that for slowly
varying forces (relative to the natural frequency of the system), the dynamic effects
are small, and damping ratio does not play an important role.

2. Depending on damping ratio, for a sufficiently larger frequency ratio, Rd is smaller


than 1 implying that the dynamic response is smaller than the static response.
When the frequency ratio goes to infinity, the dynamic response goes to zero. In
these conditions, the influence of damping is negligible.

3. If β ≈ 1, i.e., the loading frequency is close to the natural frequency of the system,
Rd is very sensitive to damping, and for smaller damping values, Rd can be several
times larger than 1, implying that the amplitude of dynamic response can be much
larger than the static response. Such a condition where the loading frequency is
equal to the natural frequency of the system is called as resonance, and is discussed
in detail in the following section.
CHAPTER 7. FORCED VIBRATION OF SDOF SYSTEMS 89

7.2.2.2 Resonance
When resonance occurs β = 1 , and in this conditions the constants A, B, C, and D can
be written as follows (see Equations 7.21, 7.22, 7.25, 7.26).

D = 0
−f0
C =
2kζ
f0
A = u0 +
2kζ
u̇0 + Aζω
B =
ωD
If the system is initially at rest, the total response can be written as

 
−ζωtf0 f0 ω f0
u(t) = e cos ωD t + sin ωD t − cos ωD t
2kζ 2kωD 2kζ
" ! #
f0 ζ
= e−ζωt
cos ωD t + p sin ωD t − cos ωD t (7.28)
2kζ 1 − ζ2

The amplitude of this motion is f0 /2kζ, which means that the static response is amplified
by an amount of 1/2ζ. If the damping ratio is small, the sinusoidal term in Equation 7.28
can be neglected and ωD ≈ ω; thus the resonant response can be written as
f0 
u(t) = cos ωt e−ζωt − 1 (7.29)
2kζ
This equation shows that the resonant response is a cosine function, and the amplitude
increases with the envelope function multiplying the cosine term in Equation 7.29. The
solution is shown in Figure 7.6. The displacement response is plotted for unit static
displacement, i.e., f0 /k = 1 m. The maximum displacement amplified 10 (1/2ζ) times.
The envelope curves which govern the increase of amplitude are shown in Figure 7.6 with
green and red lines. Note that the response does not increase indefinitely as in the case
of undamped resonance. Presence of damping bounds the resonant amplitude to a finite
value as indicated by the dashed black lines in Figure 7.6. The effect of damping when
the system is near resonance is dramatic. For illustration, the steady-state response of a
SDOF system at resonance is shown in Figure 7.7 for different levels of damping. The
undamped response at resonance increases indefinitely as indicated by the red curve in
Figure 7.7. The presence of even a small amount of damping, for example, 1% significantly
reduces the amplitude of motion. With higher damping present, the response at resonance
is further reduced.
CHAPTER 7. FORCED VIBRATION OF SDOF SYSTEMS 90

u(t)
f0
£ −ζωt ¤
2ζk e −1
−f0
£ −ζωt ¤
2ζk e −1
15

10
Displacement (m)

−5

−10

−15
0 5 10 15
Time (s)

Figure 7.6: Harmonic response of a SDOF at resonance: ζ = 0.05,f0 /k = 1 m, u0 = 0,


u̇0 = 0, T = 1 s.

ζ = 0.01
ζ = 0.05
ζ = 0.2
ζ =0
50
Displacement (m)

−50
0 5 10 15 20
Time (s)

Figure 7.7: Harmonic response of a SDOF at resonance: T = 1 s, f0 /k = 1 m, u0 = 0,


u̇0 = 0, note that the undamped solution is obtained from Equation 7.12 and the damped
solutions are obtained from Equation 7.29.
Chapter 8

Free vibration of MDOF systems

8.1 Introduction
This chapter introduces the basic theory of free vibration of MDOF systems. Starting
from the matrix differential equation of motion, the frequencies corresponding to harmonic
vibrations are derived. Unlike SDOF systems, MDOF systems possess several frequen-
cies. These frequencies can be determined by formulating and solving an appropriate
eigenvalue problem. Formulation of such an eigenvalue problem is discussed followed by
the description of modal frequencies and mode shapes. Some important properties of
mode shapes are discussed. Finally, a complete solution of the free vibration problem is
formulated by modal superposition.

8.2 Vibration frequencies and mode shapes


The dynamic equation of a MDOF free vibration problem is the following system of
differential equations (see Equation 5.17).
n o
[M ] D̈ + [K] {D} = {0} (8.1)
Assuming that the vibration is simple harmonic, the displacement vector can be expressed
as

{D} = {φ} A cos(ωt + θ) (8.2)


where {φ} represents the deflected shape of the system (which does not change with
time), ω is the frequency of free vibration, A is a scalar constant, and θis the phase angle.
Differentiating Equation 8.2 successively with respect to time gives the following
n o
D̈ = −ω 2 {φ} A cos (ωt + θ) = −ω 2 {D} (8.3)
Substituting Equations 8.2 and 8.3 in Equation 8.1 gives

−ω 2 [M ] {φ} A cos (ωt + θ) + [K] {φ} A cos(ωt + θ) = {0}


which may be written as
 
[K] − ω 2 [M ] {φ} = {0} (8.4)

91
CHAPTER 8. FREE VIBRATION OF MDOF SYSTEMS 92

This equation is called as the eigenvalue or characteristic value problem of free vibration.
The quantities ω 2 are the eigenvalues or characteristic values indicating the square of free
vibration frequencies. The corresponding eigen vectors {φ} represent the shapes of the
vibrating system and are called as mode shapes. The non-trivial solution of Equation 8.4
may be written as

det [K] − ω 2 [M ] = 0 (8.5)
and is called as the frequency equation of the system. Expanding the determinant gives
a polynomial equation in ω 2 . The order of the polynomial depends on the number of
degrees of freedom of the structure. Note that the eigenvalues are meaningful only when
the mass and stiffness matrices are non-singular, i.e., their determinants are non-zero.
Recall from Chapter 1 that the presence of rigid body motion (constrained DOFs) makes
the stiffness matrix singular. The same applies to the consistent mass matrix. Therefore
one needs to remove the equations corresponding to the constrained DOFs before solving
the eigenvalue problem. In other words, proper support conditions must be imposed,
and the dynamic problem described by Equation 8.1 is formulated for unconstrained
DOFs only. In other words, the matrices and vectors in the equation correspond to the
unconstrained degrees of freedom of the system. Therefore if there are N unconstrained
degrees of freedom, the order of the polynomial equation is N . The N roots of this
equation (ω12 , ω22 , ω32 , ω42 , . . . , ωN
2
) represent the squares of frequencies of the N modes of
vibration which are possible. The mode having the lowest frequency is called as the
first mode, the next higher frequency is the second mode, etc. It is common practice in
structural dynamics to arrange the frequencies in increasing order, i.e., ω1 < ω2 < ω3 . . . <
ωN .

8.2.1 Reduction to a standard eigenvalue problem


Finding the roots of a polynomial equation can be an uncertain process in numerical anal-
ysis, as small perturbations in the polynomial coefficients can lead to large perturbations
in the roots. Therefore, a more stable solution procedure can be outlined by reducing
Equation 8.4 to a standard eigenvalue problem, for which excellent numerical solution
algorithms are available. One way of performing such a reduction is by pre-multiplying
Equation 8.5 by [M ]−1 which gives

[M ]−1 [K] − ω 2 [I] {φ} = {0}
where [I] is an identity matrix. This equation is a standard eigenvalue problem of the
form ([A] − λ [I]) {φ} = {0} with [A] = [M ]−1 [K], and λ = ω 2 . The eigenvalues are λi
and the corresponding eivenvectors are {φi } where i = 1, √2, 3, · · · N . The positive square
roots of λi provides the vibration frequencies, i.e., ωi = λi .

8.3 Spectral and mode shape matrices


Spectral matrix is a diagonal matrix with ωi2 arranged along the principal diagonal.
CHAPTER 8. FREE VIBRATION OF MDOF SYSTEMS 93

 
ω12 0 0 · · · 0
 0 ω22 0 · · · 0 
 
 0 0 ω32 · · · 0 
[Ω] ≡   (8.6)
 .. .. .. . . 
 . . . . 0 
2
0 0 0 0 ωN
The mode shape matrix is formed by arranging the mode shapes {φi } along the columns

 
φ11 φ12 φ13 ··· φ1N
 φ21 φ22 φ23 ··· φ2N 
   
 φ31 φ32 φ33 ··· φ3N 
[Φ] = {φ1 } {φ2 } {φ3 } · · · {φN } = 
 .. .. .. .. 
 . . . ··· . 
φN 1 φN 2 φN 3 ··· φN N

where the element φij corresponds to DOF i and mode j.

8.4 Static condensation


Provided that the rigid body displacements are not permitted (the structure is sufficiently
supported), the stiffness matrix is non-singular. If all the unconstrained degrees of freedom
possess some mass (inertia) then the mass matrix is also non-singular. This is the case
when consistent mass formulation is used. In fact, the structure of the stiffness and
the consistent mass matrix is the same because they are derived from the same shape
functions. However, if lumped mass model is used with no rotational inertia at the
rotational degrees of freedom, the mass matrix becomes singular, and the eigenvalue
solution can result in imaginary frequencies. In such cases, the mass less degrees of
freedom need to be removed from the eigenvalue formulation. This can be obtained by
grouping the degrees of freedom into two categories– those that have mass are called
as dynamic degrees of freedom, and those without any mass are called as static degrees
of freedom. The corresponding components of the mode shapes are then partitioned
as dynamic component {φD } and the static component {φS }. Similarly the mass and
stiffness matrices are reordered and partitioned, and the characteristic equation is written
as

       
[KDD ] [KDS ] {φD } [MDD ] [0] {φD } {0}
−ω 2
= (8.7)
[KSD ] [KSS ] {φS } [0] [0] {φS } {0}

where [KDD ] is a sub-matrix containing rows and columns of [K]which correspond to


DOFs, [KDS ] is a sub-matrix of [K] with rows corresponding to dynamic DOFs and
columns corresponding to static DOFs, etc. Note that because [K]is symmetric, [KDS ] =
[KSD ]T . From the second partition of Equation 8.7 we obtain the following.

[KSD ] {φD } + [KSS ] {φS } = {0}


⇒ {φS } = − [KSS ]−1 [KSD ] {φD } (8.8)

From the first partition of Equation 8.7 we obtain


CHAPTER 8. FREE VIBRATION OF MDOF SYSTEMS 94

[KDD ] {φD } + [KDS ] {φS } − ω 2 [MDD ] {φD } = {0} (8.9)


Substituting Equation 8.8 in Equation 8.9 results in

[KDD ] {φD } − [KDS ] [KSS ]−1 [KSD ] {φD } − ω 2 [MDD ] {φD } = {0}

⇒ [KDD ] − [KDS ] [KSS ]−1 [KSD ] − ω 2 [MDD ] {φD } = {0}
  
⇒ K̄ − ω 2 [MDD ] {φD } = {0} (8.10)
 
where K̄ ≡ [KDD ] − [KDS ] [KSS ]−1 [KSD ] is called as the condensed stiffness matrix.
Equation 8.10 defines a reduced eigenvalue problem with as many frequencies and mode
shapes as there are dynamic degrees of freedom. Once the dynamic components of mode
shapes are determined by solving the eigenvalue problem, the static components can be
obtained from Equation 8.8. Finally the complete mode shape is obtained by combining
the static and dynamic components, i.e.,
 
{φD }
{φ} = (8.11)
{φS }

8.5 Orthogonality of mode shapes


The structural mode shapes {φi }possess certain special properties which are very useful
in dynamic analysis if structures. These properties, called as orthogonality relationships,
can be derived from Betti’s law, as described below.

Betti’s law
The work done by a set of forces 1 in going through the displacements caused by another
set of forces 2 equal to the work done by the set of forces 2 in going through the displace-
ments caused by the set of forces 1. This law leads to the orthogonality relationships of
mode shapes.
Let us consider any two modes of vibration i and j with frequencies ωi and ωj , and the
corresponding mode shapes {φi } and {φj }, respectively. The displacements corresponding
to these two modes are found from Equation 8.2

{Di } = {φi } Ai cos (ωi t + θi ) (8.12)


{Dj } = {φj } Aj cos (ωj t + θj ) (8.13)

and the corresponding accelerations are as follows.

n o
D̈i = −ωi2 {φi } Ai cos (ωi t + θi ) (8.14)
n o
D̈j = −ωj2 {φj } Aj cos (ωj t + θj ) (8.15)

The corresponding inertia forces are obtained by multiplying accelerations with the mass
matrices.
CHAPTER 8. FREE VIBRATION OF MDOF SYSTEMS 95

n o
{fi } = [M ] D̈i = −ωi2 [M ] {φi } Ai cos (ωi t + θi )
n o
{fj } = [M ] D̈j = −ωj2 [M ] {φj } Aj cos (ωj t + θj )

The work done by forces {fi }in going through displacements {Dj } is obtained as

  
Wij = {Dj }T {fi } = {φj }T Aj cos (ωj t + θj ) −ωi2 [M ] {φi } Ai cos (ωi t + θi ) (8.16)

Similarly the work done by forces {fj }in going through displacements {Di } is obtained
as

  
T T
Wji = {Di } {fj } = {φi } Ai cos (ωi t + θi ) −ωj2 [M ] {φj } Aj cos (ωj t + θj ) (8.17)

According to Betti’s law, Wij = Wji which means that Equations 8.16 and 8.17 can be
equated.

ωi2 {φj }T [M ] {φi } = ωj2 {φi }T [M ] {φj } (8.18)


Since both sides of Equation 8.18 represent scalar quantity, we can transpose the left side
to obtain

 
ωi2 φTi [M ] {φj } = ωj2 φTi [M ] {φj }
 
⇒ φTi [M ] {φj } ωi2 − ωj2 = 0 (8.19)

If the two modes i and j are different, i.e., i 6= j then ωi2 − ωj2 6= 0 which implies from
Equation 8.19 that

φTi [M ] {φj } = 0 (8.20)
This condition is said to imply that the mode shapes are orthogonal with respect to the
mass matrix. It can also be proved that the mode shapes are orthogonal with respect to
the stiffness matrix. To do so we start with Equation 8.4 for mode j.

[K] {φj } − ωj2 [M ] {φj } = {0} (8.21)


Pre-multiplying both sides by {φi }T we obtain

{φi }T [K] {φj } − ω 2 {φi }T [M ] {φj } = {φi }T {0} (8.22)


where {φi }T [M ] {φj } = 0 by virtue of Equation 8.21. Thus we have the following result.

{φi }T [K] {φj } = 0 (8.23)


This result is said to imply that the mode shapes are orthogonal with respect to the
stiffness matrix.
CHAPTER 8. FREE VIBRATION OF MDOF SYSTEMS 96

8.6 Generalized mass and generalized stiffness matrix


In the previous section we saw that any two modes i and j are orthogonal with respect
to the mass and the stiffness matrices as long as these modes have different frequencies.
If , on the other hand, the two modes in Equations 8.20 and 8.22 are the same, we obtain
scalar quantities having units of mass, and stiffness, and are called as generalized mass
and generalized stiffness. For any mode i, the generalized mass is defined as

fi = {φi }T [M ] {φj }
M (8.24)
and the generalized stiffness is defined as

e i = {φi }T [K] {φj }


K (8.25)
Then the frequency of mode i can also be obtained as
s
fi
K
ωi = (8.26)
fi
M
The orthogonality property of mode shapes implies that the matrix [Φ]T [M ] [Φ] is a
fi arranged along the main diagonal. This matrix is called as the
diagonal matrix with M
generalized mass matrix
h i
f = [Φ]T [M ] [Φ]
M (8.27)

where [Φ] is the mode shape matrix defined in Section 8.3. In a similar manner, the
generalized stiffness matrix is defined as
h i
Ke = [Φ]T [K] [Φ] (8.28)

8.7 Normalization of mode shapes


The mode shapes {φi } obtained as eigen vectors are absolute and only relative, in the
sense that their elements are found by assuming one of the elements to be equal to
1. In fact, the elements of the mode shapes can not be evaluated independently, and
are only evaluated relative to one of the elements. This means that the mode shape
scaled with any non-zero scalar α is also a mode shape of the structure. This can be
easily verified from Equation 8.4–if {φ} satisfies the equation, then α {φ} also satisfies the
equation. It implies that it is permitted, and often convenient, to scale the mode shapes
with appropriate scalars without violating the characteristic equation. There are several
methods of scaling (normalizing) the mode shapes. Some of the more popular ones are
discussed below.

Normalized to make largest term of {φ}equal to 1


In this approach, the mode shape is divided by its largest element to obtain the normalized
mode shape. The normalized mode shape has all elements between -1 and 1, and is easy
to visualize.
CHAPTER 8. FREE VIBRATION OF MDOF SYSTEMS 97

Normalized to make the Euclidean norm of {φ}equal to 1


The Euclidean norm of {φ} is {φ}T {φ}. If the original mode shape is denoted as {φo }
and the normalized mode shape is denoted as {φn }, the relation between the two is given
as
s
1
{φn } = T
{φo } (8.29)
{φo } {φo }

Normalized to make generalized mass equal to 1


In this approach, each mode shape is normalized in such a way that the generalized mass
of that mode is equal to 1. The relationship between the normalized and the original
mode shapes is given as
s
1
n
{φ } = T
{φo } (8.30)
{φ } [M ] {φ }
o o

If the system has large mass, the mode shape normalized in this manner might contain very
small numbers, which might be undesirable due to the limitation of numerical precision
of computers.

Normalized to make the first element equal to 1


In this approach the mode shape is divided by its first element. This method was popular
in old hand calculations, and were transferred to some computer programs. This kind of
normalization is not recommended because it results in serious problems if the first term
of the mode shape is close to zero.

8.8 Free vibration solution by modal expansion


Having determined the modal frequencies and mode shapes, we can apply modal expansion
to solve the free vibration problem. The motion is described by the following initial value
problem

n o n o n o
[M ] D̈ + [K] {D} = {0} , {D(t = 0)} = {D0 } , Ḋ(t = 0) = Ḋ0 (8.31)

This is a system of differential equations, which are coupled due to the non-zero off-
diagonal terms of the mass, and the stiffness matrices. This implies that each equation
of the system contains more than one displacement components which have to be solved
simultaneously. On the other hand, if the associated matrices were diagonal, the system
of equations would be uncoupled in the sense that each of the equation would involve only
one dependent variable and its derivatives. Such equations can be solved easily by using
methods described in Chapter 6. Modal expansion of displacement along with the use of
orthogonality property allows such simplification as described below.
CHAPTER 8. FREE VIBRATION OF MDOF SYSTEMS 98

8.8.1 Modal expansion of displacement


The displacement {D} is a function of both space (degree of freedom), and time. Since
the mode shapes form an independent set of vectors, they can be used to separate the
spatial and temporal variation of displacement according to the following equation.

{D} = {φ1 } y1 (t) + {φ2 } y2 (t) + . . . + {φN } yN (t) (8.32)


Here the mode shapes {φi }serve to describe the N independent displacement patterns
(shapes) while the scalar functions of time yi (t) provide the time dependent amplitudes
of motion. These amplitudes of motion represented by yi (t) are called as generalized
coordinates. This operation effectively breaks down the total displacement as a sum of
displacements due to different modes, and is therefore equivalent to superposition prin-
ciple. Such superposition is valid only for linearly elastic systems. The time dependence
of displacement in each mode is described by yi (t) and its spatial variation is described
by {φi }. The generalized coordinate yi (t) will henceforth be denoted as yi without loss
of generality. With such a notation, the summation of Equation 8.32 may be written in
matrix form as
 

 y 1 


 
  y2 
{D} = {φ1 } {φ2 } · · · {φN } .. = [Φ] {y} (8.33)


 . 

 y  
N

where [Φ] is the mode shape matrix and {y} is a collection of generalized coordinates.
This operation is called as modal expansion of displacement. These results indicate that
the mode shape matrix [Φ] can be looked upon as a transformation from generalized
coordinates to geometric coordinates. Since the mode shape matrix is independent of
time, the acceleration can be written as
n o
D̈ = [Φ] {ÿ} (8.34)
The initial displacement, and velocity can be converted to generalized coordinates by
using the following equations.

{y0 } = [Φ]−1 {D0 } (8.35)


n o
{ẏ0 } = [Φ]−1 Ḋ0 (8.36)

In practice, it is actually not necessary to invert the [Φ] matrix to find the initial conditions
in generalized coordinates. A simpler procedure can be used as outlined next. Using the
modal expansion of displacement vector we have

{D} = {φ1 } y1 + {φ2 } y2 + ... {φi } yi + ... {φN } yN (8.37)


Pre-multiplying Equation 8.37 with {φi }T [M ] we obtain

{φi }T [M ] {D} = {φi }T [M ] {φ1 } y1 + {φi }T [M ] {φ2 } y2 + ...


+ {φi }T [M ] {φi } yi + ... + {φi }T [M ] {φN } yN
CHAPTER 8. FREE VIBRATION OF MDOF SYSTEMS 99

where the orthogonal property of the mode shapes implies that all the terms on the right
hand side except for {φi }T [M ] {φi } yi vanish, thus yielding

{φi }T [M ] {D} = {φi }T [M ] {φi } yi

{φi }T [M ] {D}
⇒ yi = (8.38)
fi
M
This equation provides the transformation of the structural coordinates (also called ge-
ometrical coordinates) to the generalized coordinates. Applying this at the initial time,
gives us the initial conditions in generalized coordinates, as

{φi }T [M ] {D0 }
y0i = (8.39)
fi
M
n o
{φi }T [M ] Ḋ0
ẏ0i = (8.40)
fi
M

8.8.2 Modal decoupling of the equation of motion


By using the modal expansion concept, the coupled system of Equation 8.31 can be
decoupled. Substituting Equations 8.33 and 8.34 in Equation 8.31 we obtain

[M ] [Φ] {ÿ} + [K] [Φ] {y} = {0}


which, after pre-multiplication by [Φ]T results in the following equation.

[Φ]T [M ] [Φ] {ÿ} + [Φ]T [K] [Φ] {y} = {0} (8.41)


Using the generalized mass and generalized stiffness matrices defined in Section 8.6 we
get the following.
h i h i
Mf {ÿ} + K e {y} = {0} (8.42)
Because the generalized matrices are diagonal, the of equations is uncoupled, and each
of qi is given by the following second order homogeneous differential equation

fi ÿi + K
M e i yi = 0 (8.43)
with initial conditions y0i and ẏ0i obtained from Equations 8.39 and 8.40. This equa-
tion can be solved by using standard methods described in Chapter 6. The solution is
reproduced here for convenience.
ẏ0i
yi = y0i cos ωi t +sin ωi t (8.44)
ωi
Once, all the generalized coordinates yi i = 1, 2, 3, · · · N have been determined,
they can be transformed according to Equation 8.33 to obtain {D}.
CHAPTER 8. FREE VIBRATION OF MDOF SYSTEMS 100

8.9 Damping models


In real structures, vibration energy is dissipated by mechanisms called as damping. With
analogy
n o to SDOF systems, viscous damping force nin o a structure can be represented as
[C] Ḋ where [C] is the damping matrix, and Ḋ is a vector of velocities at the
selected DOFs. It is impractical, and often impossible, to determined the coefficients
of the damping matrix directly from structural dimensions, structural member size, and
the damping properties of the structural materials used. It is a common practice to
directly specify damping ratios is different modes of vibration based on past experience
and experiments conducted on different structural types. There are several empirical
models of creating viscous (or classical) damping matrices. Some of the more popular
models are described in this section.

8.9.1 Classically damped systems


Classical damping models assume that energy dissipation mechanisms are uniformly dis-
tributed throughout the structure, for example a multistory building with a similar struc-
tural system and structural material over its height. In these cases, it is reasonable to
assume that damping is contributed by the structural elements which also contribute to
its stiffness and mass. Based on this argument, damping matrix can be assumed to be
proportional to the stiffness matrix, the mass matrix, or both of them. Such models of
damping are called as classical damping models.

8.9.1.1 Mass proportional damping


These models assume that the damping matrix is proportional to the mass matrix, i.e.,

[C] = a0 [M ] (8.45)
with a0 as the proportionality constant to be determined. Pre-multiplying by [Φ]T and
post-multiplying by [Φ] we obtain

[Φ]T [C] [Φ] = a0 [Φ]T [M ] [Φ]


h i h i
⇒ C e = a0 M f (8.46)
h i
where C e ≡ [Φ]T [C] [Φ] is called as the generalized damping matrix. Since the general-
ized mass matrix is diagonal, it can be easily seen from Equation 8.46 that the generalized
damping matrix is also diagonal. The generalized damping coefficient of any mode i can
be written as

ei = a0 M
C fi (8.47)
ei = 2ζi M
and if ζi is the specified damping ratio of mode i then we have C fi ωi , which
results in
a0
ζi = (8.48)
2ωi
Figure 8.1 shows the variation of modal damping ratio as a function of frequency according
to Equation 8.48. The damping ratio decreases inversely with the frequency. This model
CHAPTER 8. FREE VIBRATION OF MDOF SYSTEMS 101

0.25

Damping ratio, ζ
0.2

0.15

0.1

0.05

0
0 20 40 60 80 100
Angular frequency, ω (rad/s)

Figure 8.1: Damping ratio as a function of frequency for mass proportional model

allocates higher damping ratios to the lower modes and smaller damping ratios for higher
modes. If a structures has modes spanning a wide range of frequencies, this model might
allocate inappropriate damping values to some of the modes, which can result in serious
distortion of response contribution to those modes.

8.9.1.2 Stiffness proportional damping


These models assume that the damping matrix is proportional to the stiffness matrix, i.e.,

[C] = a1 [K] (8.49)


Where a1 is the stiffness proportional coefficient to be determined and has units of sec.
Pre-multiplying by [Φ]T and post-multiplying by [Φ] we obtain

[Φ]T [C] [Φ] = a1 [Φ]T [K] [Φ]


h i h i
e e
C = a1 K
This shows that the generalized damping matrix is diagonal. For any mode i, the damping
coefficient can be found as

ei = a1 K
C ei
ei = 2ζi M
and if ζi is the specified value of damping ratio of mode i then we have C fi ωi ,
which give us

fi ωi = a1 K
2ζi M ei

ωi
ζi = a1 (8.50)
2
CHAPTER 8. FREE VIBRATION OF MDOF SYSTEMS 102

0.09

0.08

0.07

Damping ratio, ζ
0.06

0.05

0.04

0.03

0.02

0.01

0
0 20 40 60 80 100
Angular frequency, ω (rad/s)

Figure 8.2: Damping ratio as a function of frequency for stiffness proportional model

Figure 8.2 shows the variation of modal damping ratio as a function of frequency according
to Equation 8.50. The damping ratio increases linearly with the frequency. This model
allocates larger damping ratios to the higher modes and smaller damping ratios for lower
modes. If a structures has modes spanning a wide range of frequencies, this model might
allocate inappropriate damping values to some of the modes, which can result in serious
distortion of response contribution to those modes.

8.9.1.3 Rayleigh damping


The drawbacks of mass and stiffness proportional damping can be overcome to some
extent by using a damping model which is proportional to both the mass and the stiffness
matrices. Such a damping model is called as Rayleigh damping, where the damping
matrix is given as

[C] = a0 [M ] + a1 [K] (8.51)


Using similar procedure as above, the damping ratio of mode i is obtained as
a0 a1 ωi
+ ζi = (8.52)
2ωi 2
There are two unknown coefficients in this equation, which can be determined by speci-
fying the damping ratios and frequencies of any two modes, say p and q, as ζp , ζq , ωp , ωq .
Using these values successively in Equation 8.52 gives the following two equations
a0 a1 ωp
ζp = +
2ωp 2
a0 a1 ωq
ζq = +
2ωq 2
CHAPTER 8. FREE VIBRATION OF MDOF SYSTEMS 103

Mass proportional
Stiffness proportional
0.25 Rayleigh

Damping ratio, ζ
0.2

0.15

0.1

0.05

0
0 20 40 60 80 100
Angular frequency, ω (rad/s)

Figure 8.3: Damping ratio as a function of frequency

which can be simultaneously solved to obtain the proportionality constants


2ωp ωq
a0 = (ζq ωp − ζp ωq )
ωp2 − ωq2
2
a1 = (ζp ωp − ζq ωq )
ωp2 − ωq2
If the damping ratios in the two modes are taken as equal, i.e., ζm = ζn = ζ, then the
equations can be further simplified as
2ωp ωq
a0 = ζ (8.53)
ωp + ωq
2
a0 = ζ (8.54)
ωm + ωn
The variation of damping ratio with frequency according to Equation 8.52 is shown
in Figure 8.3. The damping model is constructed to have 5% of critical damping at two
modes with frequencies 10 rad/s and 50 rad/s.

8.9.1.4 Damping matrix by superposition of modal damping ratios


If it is desirable to specify the damping ratios of all the modes of vibration, then the
construction of damping matrix is straightforward, and is discussed in this section. Let
the damping ratios of all the modes be specified, a typical mode having a damping ratio ζi .
Then the damping coefficients of the mode is simply given as C ei = 2ζi M
fi ωi . A diagonal
CHAPTER 8. FREE VIBRATION OF MDOF SYSTEMS 104

h i
matrix containing these damping coefficients gives the generalized damping matrix C e
h i
which is related to the damping matrix by the equation C e = [Φ]T [C] [Φ], from which
we get
 −1 h i
[C] = [Φ] T e [Φ]−1
C (8.55)
Since matrix inversion is computationally expensive, this equation can be equivalently
computed as
!
XN
2ζi ωi
[C] = [M ] {φi } {φi }T [M ] (8.56)
f
Mi
i=1

8.9.2 Non-classically damped systems


For structures made up of more than a single type of material, where the different materials
provide drastically different energy-loss mechanisms in various parts of the structure, the
distribution of damping forces will not be similar to the distribution of the inertial and
elastic forces. An example of such a structure is a soil-structure system. The modal
damping ratio for the soil system would be typically much different than that of the
structure, usually 3 to 4 times larger. Another example would include structural systems
equipped with supplemental damping mechanisms that are concentrated equipments that
provide substantially large damping than the structure itself. In these situations the
damping matrix will not be proportional to the mass and the stiffness matrices. Such
systems are called non-classically damped systems, and the resulting damping matrix is
called as non-classical or non-proportional damping matrix. Mode shapes of free vibration
are, in general, not orthogonal with respect to the non-proportional damping matrices.
This implies that the modal solution method is not feasible in these situations even when
the system is linearly elastic. Direct integration of the matrix equation of motion using
numerical algorithms such as Newmark’s method provides a feasible solution of non-
classically damped systems.
The construction of damping matrix for these types of systems uses the approximation
that the damping matrix for each sub-system is proportional. In other words, the whole
system is divided into different sub-structures, each of which has a different damping
mechanism, that is modeled as classical damping as discussed in the previous sections.
Finally the damping matrices of the sub-systems are combined together to obtain the
damping matrix of the whole structure.

8.10 Free vibration of classically damped MDOF sys-


tems
If [C]represents n
the o
classical damping matrix, free vibration problem with initial condi-
tions {D0 } and Ḋ0 is described by the following equation

n o n o n o n o
[M ] D̈ +[C] Ḋ +[K] {D} = {0} , {D(t = 0)} = {D0 } , D(t ˙= 0) = Ḋ0 (8.57)
CHAPTER 8. FREE VIBRATION OF MDOF SYSTEMS 105

n o n o
By using the modal expansion, {D} = [Φ] {y}, Ḋ = [Φ] {ẏ} and D̈ = [Φ] {ÿ} , the
equation is converted to generalized coordinates as

[M ] [Φ] {ÿ} + [C] [Φ] {ẏ} + [K] [Φ] {y} = {0} (8.58)
This coupled system of equation can be uncoupled by pre-multiplying with [Φ] which T

gives
h i h i h i
f e
M {ÿ} + C {ẏ} + K e {y} = {0} (8.59)
Due to the diagonal nature of the generalized matrices, this equation gives N independent
equations of the form

fi ÿi + C
M ei ẏi + K
eiy = 0 (8.60)
This is equivalent to damped free vibration of a SDOF system with mass M fi , damping
ei , stiffness K
coefficient C e i , initial displacement y0i and initial velocity ẏ0i (see Equations
8.39 and 8.40), the solution of which can be obtained from methods described in Chapter
6. The solution is reproduced below.
 
ẏ0i + ζi ωi y0i
yi = y0i cos ωDi t + sin ωDi (8.61)
ωDi
p
where ζi ≡ C ei /(2M fi ωi ) is the damped frequency of mode i, and ωDi ≡ ωi 1 − ζ 2 is the
i
damped frequency of mode i. Once all the yi , i = 1, 2, 3, · · · N have been computed, they
are arranged in a vector to form {y} which is transformed to geometrical coordinates to
obtain {D} = [Φ] {y}.

You might also like