Lecture Notes in Computational Mechanics
Lecture Notes in Computational Mechanics
Lecture notes
1
CONTENTS 2
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.
4
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 5
B A P
(a)
Δ k
1
B A Δ
P
(b) (c)
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.
k
u1, f1 u2, f2
1 2
f1 + f2 = 0
⇒ f1 = −f2 (1.1)
CHAPTER 1. INTRODUCTION TO THE STIFFNESS METHOD 6
f2 = k(u2 − u1 )
⇒ f2 = −ku1 + ku2 (1.2)
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.
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.
k1 k2
P Q
1 2 3
k1
f1, u1
f21, u2
1 2
k2
f22, u2
f3, u3
1 2
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
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 ( )
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
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
1.3.1 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)
∂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
−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
A1 E A1 L
F1 = U1 − U2 + (0) U3 = −P
L E
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.
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)
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.
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
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
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, ∆
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)
∂Πp
= 0
∂u1
⇒ k1 U1 − k1 U2 − F1 = 0
F1 = k1 U1 − k1 U2 + 0 (U3 ) (1.21)
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.
2 3
1 2 1m
1 4
3 X
1m
1kN
x
D2j
dj
D2j-1
j(Xj,Yj)
D2i
y
α
Y D2i-1
di i(Xi,Yi)
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
∂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
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
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
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
F7= 207
F3= 0
1 2
F8= 0
F7= -207
3
F1= 0
F2= -1000
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
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.
27
CHAPTER 2. FINITE ELEMENT FORMULATION OF ELASTIC CONTINUUM 28
T
P
ST
B
Y
SU
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 }.
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
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)
Eν
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.
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
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
W = WˆT + W B ˆ
T
= − {U } {B} dV − {U }T {T } dS (2.9)
V ST
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.
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.
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
h 0i
[k] = [T ]T k [T ]
n 0o
{f } = [T ]−1 f
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
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
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
2.13 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
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.
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.
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
D1 L
Y
d1 1
X
Figure 3.2: A truss element: local and global DOFs
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
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
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
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.
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 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
n 0o
{e} = {exx } = [S] d
1 0.88
= 1 −1 × 10−3
L 0
= 3.93 × 10−5
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.
neutral surface
compression
tension
We consider beams have a straight longitudinal axis and are prismatic (cross section
is constant).
50
CHAPTER 4. DEVELOPMENT OF BEAM MEMBERS 51
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.
• 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
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)
θ
θ
B'
y
A'
uy(x,y = 0)
B
y
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
d1 d3
d4
d2
x
1 2
With the 4 DOF available, a cubic polynomial can be used to interpolate the displacement
field within the element, i.e.,
−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
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.
d1=1 d2=1
d3=d4=0 d3=d4=0
d1=0
d3=1
d1=d2=0 d1=d2=0 d3=0
d4=1
∂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)
[E] = E (4.15)
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
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
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
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.
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
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
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
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
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 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.
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
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.
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
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
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 evaluatedsimilarly, 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
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.
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)
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
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.
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.
u(t)
k
m f(t)
c
Figure 6.3: A mechanical model of a SDOF dynamic system with viscous damping.
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
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
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.
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.
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.
The arbitrary constants A1 + A2 and A1 − A2 are denoted as constants A and B, and the
solution written 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)
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)
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.
The complex exponential in the above equation can be converted to trigonometric func-
tions by suing the Euler’s equations.
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.
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)
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.
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
can be obtained as
δ
ζ≈ (6.48)
2π
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
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.
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
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.
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
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)
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.
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.
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)
√
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.
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)
ζ = 0.01
ζ = 0.05
ζ = 0.2
ζ =0
50
Displacement (m)
−50
0 5 10 15 20
Time (s)
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.
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 .
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
[KDD ] [KDS ] {φD } [MDD ] [0] {φD } {0}
−ω 2
= (8.7)
[KSD ] [KSS ] {φS } [0] [0] {φS } {0}
[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 }
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
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 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.
fi = {φi }T [M ] {φj }
M (8.24)
and the generalized stiffness is defined as
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)
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.
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
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.
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
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}
⇒ 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
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
[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
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.
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.
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)
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
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}.