Fem
Fem
Fem
Dr. C. S. Jog
Dept. of Mechanical Engineering
Indian Institute of Science
Bangalore-560012
Chapter 1
Calculus of Variations
1.1 Introduction
In 1696, the Swiss mathematician Johann Bernoulli challenged his colleagues
to solve an unresolved issue called the brachistochrone problem which is as
follows: Find the shape of a frictionless chute between points (1) and (2) in
a vertical plane such that a body sliding under the action of its own weight
goes from point (1) to point (2) in the shortest time. Bernoulli originally
specified a deadline of six months, but extended it to a year and half at the
request of Leibniz, one of the leading scholars of the time, and the man who
had, independently of Newton, invented the differential and integral calculus.
The challenge was delivered to Newton at four p.m. on January 29, 1697.
Before leaving for work the next morning, he had invented an entire new
branch of mathematics called the calculus of variations, used it to solve the
brachistochrone problem and sent off the solution, which was published at
Newtons request, anonymously. But the brilliance and originality of the
work betrayed the identity of its author. When Bernoulli saw the solution,
he commented, We recognize the lion by his claw. Newton was then fifty-
five years old. Much of the formulation of this discipline was also developed
by the Swiss mathematician Leonhard Euler.
What is the relevance of the calculus of variations to the finite element
method? For one, the finite element formulation can be derived in a more
direct way from a variational principle than from the corresponding energy
functional. But perhaps more importantly, there are several problems such
as, for example, inelastic deformations, where the governing differential equa-
tions can be cast in variational form but for which no corresponding energy
functional exists. In such cases, developing the finite element formulation
from the variational principle is the only alternative. We start by presenting
the principles of the calculus of variations. The treatment presented here
follows closely the one given in Shames and Dym [2].
1
1.2 Principles of the Calculus of Variations
While dealing with a function y = f (x) in the ordinary calculus, we often
want to determine the values of x for which y is a maximum or minimum. To
establish the conditions of local extrema, we expand f (x) as a Taylor series
about a position x = a:
1 d2 f
df
f (x) = f (a) + (x a) + 2
(x a)2 + . . . , (1.1)
dx x=a
2! dx x=a
or alternatively,
1 d2 f
df
f (x) f (a) = x + (x)2 + . . . ,
dx x=a 2! dx2 x=a
where x = x a. Since x can be positive or negative, the necessary
condition for a local maxima or minima at an interior point (i.e., point not
lying on the boundary of the domain) x = a is
df
= 0.
dx x=a
The point x = a is called an extremal point.
In the calculus of variations, we are concerned with extremizing (mini-
mizing or maximizing) functionals which, roughly speaking, are functions of
functions. An example of a functional is
Z x2
I= F (x, y, y )dx,
x1
2
1 x
y=f(x)
y
g 2
3
y
1 ~y(x)
y(x)
~y(x)
2
Figure 1.2: Extremal path and the varied paths.
where F is a known function. Assume that y(x1 ) and y(x2 ) are given (e.g.,
the starting and ending points in the brachistochrone problem). Let y(x) be
the optimal path which extremizes (minimizes or maximizes) I, and let y(x)
be varied paths as shown in Fig. 1.2. Note that y(x) agrees with y(x) at the
end points y(x1 ) and y(x2 ) since these are known. Our goal is to expand the
functional I in the form of a Taylor series similar to Eqn. 1.1 so that the usual
principles of calculus can be used for obtaining the path which minimizes or
maximizes I. Towards this end, we relate y(x) and y(x) by using a scalar
parameter as follows:
y(x) = y(x) + (x), (1.8)
where (x) is a differentiable function having the requirement (x1 ) = (x2 ) =
0, so that y(x1 ) = y(x1 ) and y(x2 ) = y(x2 ). We see that an infinite number
of paths can be generated by varying . In terms of the varied paths y(x),
the functional in Eqn. 1.7 can be written as
Z x2 Z x2
I=
F (x, y, y )dx = F (x, y + , y + )dx. (1.9)
x1 x1
4
, we can expand it as a power series similar to the one in Eqn. 1.1. Thus,
d
I d 2
I 2
I = I + + + ..., (1.10)
d d2 2!
=0
=0 =0
or, alternatively,
d
I d2 I 2
I I = I I = + 2 + ....
d d 2!
=0
=0 =0
or in other words,
x2
F y F y
Z
+ dx = 0.
x1 y y =0
5
Thus, y is simply the vertical distance between points on different curves at
the same value of x. Since
dy dy dy d d(y)
= = (y y) = ,
dx dx dx dx dx
or, equivalently,
F
F
y + y +o (y)2 +o (y )2 .
F (x, y +y, y +y )F (x, y, y ) =
y y
(1.13)
The left hand side of the above equation is the total variation of F , which
we denote as (T ) F . We call the bracketed term on the right-hand side of the
equation as the first variation, (1) F . Thus,
(T ) I = I I,
6
x2
F F
Z
(1)
I= y + y dx.
x1 y y
1.2.4 Examples
In this subsection, we shall solve the brachistochrone and geodesic problems
which were presented earlier.
The Brachistochrone Problem:
We have seen that the functional for this problem is given by Eqn. 1.2. We
thus have
Z x2 s
1 1 + (y )2
I= dx,
2g x1 y
or in other words s
1 + (y )2
F = .
y
Substituting for F in the Euler-Lagrange equation given by Eqn. 1.11, we
get the differential equation
2yy + 1 + (y )2 = 0.
7
To solve the above differential equation, we make the substitution y = u to
get
du 1 + u2
u = ,
dy 2y
or
udu dy
Z Z
2
= .
1+u 2y
Integrating, we get
y(1 + u2 ) = c1 ,
or, alternatively
y(1 + (y )2 ) = c1 .
The above equation can be written as
Z
ydy
x= + x0 .
c1 y
x = R sin cos ,
y = R sin sin , (1.17)
z = R cos .
8
we have " 2 #1/2
d
F = R 1 + sin2 .
d
Noting that F is a function of but not of , the Euler-Lagrange equation
when integrated gives the solution
cos c1 sin sin sin c1 sin cos = c2 cos ,
where c1 and c2 are constants. Using Eqn. 1.17, the above equation can also
be written as
(cos c1 )y (cos c1 )x = c2 z,
which is nothing but the equation of a plane surface going through the origin.
Thus, given any two points on the sphere, the intersection of the plane passing
through the two points and the origin, with the sphere gives the curve of the
shortest length amongst all curves lying on the surface of the sphere and
joining the two points. Such a curve is known as a Great Circle.
In this example, the constraint that the curve lie on the surface of the
sphere was treated implicitly by using spherical coordinates, and thus did
not have to be considered explicitly. This procedure might not be possible
for all problems where constraints on functions are involved. In such cases,
we have to devise a procedure for handling constraints, both of the functional
and function type. We shall do so in Sections 1.2.6 and 1.2.7.
Setting the first variation of the above functional to zero, i.e., (1) I = 0, we
get
Z x2
F F F F
y1 + + yn + y1 + + yn dx = 0.
x1 y1 yn y1 yn
Since the variations can be arbitrary, choose y2 = y3 = . . . = yn = 0.
Then since y1 is arbitrary, we get the first Euler-Lagrange equation as
F d F
= 0.
y1 dx y1
Similarly, by taking y2 as the only nonzero variation, we get the second
Euler-Lagrange equation with the subscript 1 replaced by 2. In all, we get n
Euler-Lagrange equations corresponding to the n dependent variables:
F d F
= 0, i = 1, 2, . . . , n. (1.18)
yi dx yi
9
x1 x2
k1 k2 k1
M M
To illustrate the above formulation, consider the problem shown in Fig. 1.3.
Our goal is to write the equations of motion of the two mass particles. In
order to do that, we use Hamiltons principle (see Section 6.2) which states
that for a system of particles acted on by conservative forces, the paths taken
from a configuration at time t1 to a configuration at time t2 are those that
extremize the following functional:
Z t2
I= (T V ) dt,
t1
where T is the kinetic energy of the system, while V is the potential energy
of the system. For the problem under consideration, we have
1 1
T = M x21 + M x22 ,
2 2
1 1 1
V = k1 x21 + k2 (x2 x1 )2 + k1 x22 .
2 2 2
Thus,
t2
1 1 1 2
Z
M x21 + M x22 k1 x1 + k2 (x2 x1 )2 + k1 x22
I= dt.
t1 2 2 2
Note that x1 , x2 are the two dependent variables, and t is the independent
variable in the above equation. From Eqn. 1.18, we have
F d F
= 0,
x1 dt x1
F d F
= 0.
x2 dt x2
Substituting for F , we get
M x1 + k1 x1 k2 (x2 x1 ) = 0,
M x2 + k1 x2 + k2 (x2 x1 ) = 0,
which are the same as Newtons laws applied to the two masses. By inte-
grating the above equations using the initial conditions x1 (0), x2 (0), x1 (0),
x2 (0), we can find the subsequent motions of the masses x1 (t) and x2 (t).
In this example, we could have more easily employed Newtons laws di-
rectly. However, there are many problems where it is easier to proceed by the
variational approach to arrive at the governing equations of motion. In addi-
tion, variational formulations also yield the appropriate boundary conditions
to be imposed as we shall see in future sections.
10
1.2.6 Functional constraint
In the isoperimetric problem, we saw that we have to minimize a functional
subject to a constraint on a functional. So far, we have dealt only with the
minimization of unconstrained functionals. We now show how to carry out
the extremization process in the presence of a functional constraint.
Consider the problem of minimizing
Z x2
I= F (x, y, y ) dx,
x1
y(x) = y(x) + y1 + y2 .
The variation y1 is arbitrary, while the variation y2 is such that Eqn. (1.19)
is satisfied. For extremizing I, we set (1) I to zero, while (1) J is zero since
J is a constant. Hence, using the method of Lagrange multipliers, we have
(1) I + (1) J = 0,
11
then we need to use an n + 1-parameter family of varied paths
y = y + y1 + y2 + + yn+1 ,
and n constraints
Z x2
Gk (x, y1 , y2, . . . , yp , y1 , y2 , . . . , yp ) dx = ck , k = 1, 2, . . . , n,
x1
12
x d x y
+ = 0.
2 d 2 (x2 + y 2 )1/2
(x c2 ) dx + (y c1 ) dy = 0.
(x c2 )2 + (y c1 )2 = c23 ,
which is the equation of a circle with center at (c2 , c1 ) and radius c3 . Thus, as
expected, the curve which encloses the maximum area for a given perimeter
is a circle.
Gk (x, y1 , y2 . . . , yn , y1 , y2 , . . . , yn ) = 0, k = 1, 2, . . . , m.
F d F
= 0, i = 1, 2, . . . , n,
yi dx yi
where F = F + m
P
k=1 k (x)Gk . Note that the Lagrange multipliers, k (x),
are functions of x, and are not constants as in the case of functional con-
straints.
13
1.2.8 A note on boundary conditions
So far, while extremizing I, we assumed that the value of the extremizing
function y(x) was specified at the endpoints x1 and x2 . This resulted in
y(x1 ) = y(x2 ) = 0, and hence to the elimination of the boundary terms
in Eqn. 1.16. Such a boundary condition where y(x) is specified at a point
is known as a kinematic boundary condition. The question naturally arises:
Are there other types of boundary conditions which are also valid? The
answer is that there are.
Let Z x2
I= F (x, y, y ) dx g0 y|x=x2 ,
x1
where g0 is a constant. Rewrite using Eqn. 1.16 the necessary condition for
extremizing the functional I as
Z x2
(1) F d F F F
I= y dx+ g0 y y = 0 y.
x1 y dx y y x2 y x1
(1.21)
Since the variations y are arbitrary, we can choose variations y such that
y|x=x1 = y|x=x2 = 0.
Then we get
x2
F d F
Z
y dx = 0,
x1 y dx y
for all y which satisfy y|x=x1 = y|x=x2 = 0. This in turn leads us to the
Euler-Lagrange equations
F d F
= 0.
y dx y
Hence, Eqn. 1.21 can be written as
F F
g0 y y = 0 y.
y x2 y x1
Consider variations such that y|x=x2 = 0. From the above equation, we get
F
y = 0,
y x1
14
Similar to the above process, by considering variations such that y|x=x1 = 0,
we get the boundary conditions at x = x2 as
F
= g0 or y prescribed at x2 .
y x=x2
Summarizing, the variational process has resulted in not only yielding the
governing differential equation but also the boundary conditions which can
be specified. We can either specify
1. kinematic boundary conditions at both endpoints,
For finding the extremizing function y(x), we set the first variation of I to
zero, i.e.,
Z x2
(1) F F F F
I= y + y + y + y dxM0 y |x=x2 + V y |x=x2 = 0.
x1 y y y y
15
Since the variations, y, are arbitrary, the above equation yields the Euler-
Lagrange equation
d3 d2
F F d F F
+ = 0,
dx3 y dx2 y dx y y
In each row, the first condition corresponds to the kinematic boundary con-
dition, while the second one corresponds to the natural boundary condition.
We shall consider the extension of the above concepts to more than one in-
dependent variable directly by dealing with problems in elasticity and heat
transfer in three-dimensional space. Before we do that, we consider some
basic concepts in functional analysis. We shall need these concepts in formu-
lating the variational principles and in presenting the abstract formulation.
a(u, v) = a(v, u) u, v V.
16
Now we define some Hilbert spaces which arise naturally in variational
formulations. If I = (a, b) is an interval, then the space of square-integrable
functions is Z b
L2 (I) = v: v 2 dx < .
a
17
Z 1/2
2
kvkH 1 () = (v + v v) d ,
H01 () = {v H1 () : v = 0 on } .
+ b = 0 on , (1.22)
= C( 0 ) + 0 on , (1.23)
1
(u) + (u)t on ,
= (1.24)
2
t = n on , (1.25)
t = t on t , (1.26)
18
u = u on u , (1.27)
where 0 and 0 are the initial strain and initial stress tensors, n is the out-
ward normal to , and t are the prescribed tractions on t . In a single-field
displacement-based variational formulation which we shall describe shortly,
we enforce Eqns. 1.23-1.25, and Eqn. 1.27 in a strong sense, i.e., at each
point in the domain, while the equilibrium equation given by Eqn. 1.22, and
the prescribed traction boundary condition given by Eqn. 1.26 (which is the
natural boundary condition) are enforced weakly or in an integral sense. For
the subsequent development, we find it useful to define the following function
spaces:
Vu = {v H 1 () : v = 0 on u },
Lu = u + Vu where u H 1 () satisfies u = u on u .
In what follows, we shall use direct tensorial notation. The corresponding
formulation in indicial notation is given in [2].
Now applying the divergence theorem to the first term of the above equation,
we get
Z Z Z
t
( v) n d [v : v b] d + v (t t) d = 0 v Vu .
t
(1.30)
Using Eqn. 1.25 and the fact that v = 0 on u , we have
Z Z Z
t
( v) n d = v t d = v t d v Vu .
t
19
Using the symmetry of the stress tensor, the above equation can be written
as
1
Z Z Z
t
v + (v) : d = v b d + v t d v Vu .
2 t
Now using the divergence theorem, and the fact that v = 0 on u , we get
Z Z
v ( + b) d + v (t n) d = 0 v Vu . (1.32)
t
Since the variations v are arbitrary, assume them to be zero on t , but not
zero inside the body. Then
Z
v ( + b) d = 0,
which in turn implies that Eqn. 1.22 holds. Now Eqn. 1.32 reduces to
Z
v (t n) d = 0 v Vu .
t
20
Again using the fact that the variations are arbitrary yields Eqn. 1.25.
Note that the principle of virtual work is valid for any arbitrary consti-
tutive relation since we have not made use of Eqn. 1.23 so far. Hence, we
can use the principle of virtual work as stated above for nonlinear elasticity,
plasticity etc., provided the strains are small (since we have used Eqn. 1.24
which is valid only for small displacement gradients). The variational state-
ment for a linear elastic material is obtained by substituting Eqn. 1.23 in
Eqn. 1.31. We get the corresponding problem statement as
Find u Lu such that
Z Z
(v) : C(u) d = v t d +
t
Z
v b + (v) : C0 (v) : 0 d
v Vu . (1.33)
For elastic bodies, we can formulate the method of total potential energy
which is equivalent to the variational formulation already presented as we
now show
= (1) U d v Vu ,
where we have used the fact that the delta-operator and integral commute.
Using the same fact on the left hand side of the above equation, we get
Z Z
(1)
b u d + t u d = (1) U.
t
21
or, alternatively,
(1) [U V ] = (1) = 0,
where is the potential energy defined as
= U V, (1.35)
with
Z
U= U d,
Z Z
V = b u d + t u d.
t
We have shown that the principle of virtual work is equivalent to the van-
ishing of the first variation of the potential energy. We now show that the
converse is also true. i.e., (1) = 0 implies the principle of virtual work.
Thus, assume that (1) = (1) [U V ] = 0 holds, or alternatively,
Z Z Z
(1)
U d b v d t v d = 0 v Vu .
t
22
1.5.1 Strong form of the governing equations
Let be an open domain whose boundary is composed of two open,
disjoint regions, = T q . The governing equations are
q = Q on ,
q = kT on ,
T = T on T ,
h(T T ) + (kT ) n = q on q ,
where T is the absolute temperature, q is the heat flux, Q is the heat gen-
erated per unit volume, k is a symmetric second-order tensor known as the
thermal conductivity tensor, n is the outward normal to , T is the ab-
solute temperature of the surrounding medium (also known as the ambient
temperature), T is the prescribed temperature on T , and q is the externally
applied heat flux on q . Note that q is taken as positive when heat is supplied
to the body. Typical units of the field variables and material constants in
the SI system are
T : o C, kij : W/(m-o C), Q : W/m3 , q : W/m2 , h : W/(m2 -o C).
Applying the divergence theorem to the first term on the left hand side, and
noting that v = 0 on T , we get
Z Z
[v (kT ) vQ] d + v [q + h(T T )] d = 0 v VT .
q
23
Taking the known terms to the right hand side, we can write the variational
statement as
Find T LT such that
Z Z Z Z
v (kT ) d + vhT d = vQ d + v(q + hT ) d v VT .
q q
(1.37)
We have shown how the variational form can be obtained from the strong
form of the governing equations. By reversing the steps in the above proof,
we can prove the converse, i.e., the governing differential equations can be
obtained starting from the variational formulation, provided the solution has
sufficient regularity. The energy functional corresponding to the weak form
is
Z
1 1 2
Z
= [T (kT ) QT ] d + hT hT T qT d. (1.38)
2 q 2
One can easily verify that (1) = 0 yields the variational formulation, and
vice versa. By using the positive-definiteness of k, we can show that the vari-
ational formulation is equivalent to minimizing the above energy functional.
We have seen that it is possible to give an energy formulation which is
equivalent to the variational formulation, both in the case of the elasticity
and heat transfer problems. So the following questions arise
1. Is it possible to provide an abstract framework for representing prob-
lems in different areas such as elasticity and heat transfer?
2. Can one find the appropriate energy functional from the variational
statement in a systematic way?
The answer to these questions is that an abstract formulation is possible for
self-adjoint and positive-definite bilinear operators. The elasticity and heat
transfer problems already considered are examples which fall under this cat-
egory. There are several other problems from fluid mechanics and elasticity
which also fall in this category. An abstract formulation provides a unified
treatment of such apparently diverse problems. With the help of the abstract
formulation, one can not only derive the energy functional in a systematic
way, but also prove uniqueness of solutions. In addition, it is also possible
to give an abstract formulation for the finite element method based on such
an abstract formulation which makes it possible to derive valuable informa-
tion such as error estimates without having to treat each of the problems,
whether they be from elasticity, fluid mechanics or heat transfer, individually.
We now discuss the details of such a formulation.
24
expressed in the form
Problem (V): Find u such that
a(u, v) = L(v) v V
where a(., .) is a bilinear operator, and L(.) is a linear operator with the
following properties:
1. a(., .) is symmetric (or self-adjoint).
2. a(., .) is continuous, i.e.,
L(v) kvk .
where
1
F (v) = a(v, v) L(v).
2
We have the following result relating problems (V) and (M):
Theorem 1.6.1. The minimization problem (M) is equivalent to problem
(V). The solution to Problems (M) and (V) is unique.
Proof. First we prove that M = V. Let v V and . Since u is a
minimum,
F (u) F (u + v) , v V.
Using the notation g() F (u + v), we have
g(0) g() ,
1 2
F (v) = a(u, u) L(u) + [a(u, v) L(v)] + a(v, v).
2 2
Since a(u, v) = L(v), the above equation reduces to
2
F (v) = F (u) + a(v, v) F (u).
2
Noting that F (v) = F (u) when = 0, we can write the above equation as
a(u1 , v) = L(v) v V,
a(u2 , v) = L(v) v V.
a(u1 u2 , v) = 0 v V.
Choosing v = u1 u2 , we get
a(u1 u2 , u1 u2 ) = ku1 u2 ka = 0.
ku1 u2 kV = 0,
d2 d2 v
EI 2 q = 0,
dx2 dx
d2 v d2 v
dv d
v(0) = = 0; EI = M0 ; EI 2 = V.
dx x=0 dx2 x=L dx dx x=L
26
The variational form is obtained as follows:
First define the appropriate function space in which v lies in:
Vu = H 2 (0, L), = 0, = 0 at x = 0 ,
where
L L L
d2 v d2 d2 v 2
d d v d
Z
EI 2 2 dx + EI 2 EI 2
0 dx dx dx dx dx dx 0
2
0 2
Z L
d v d d dv
+ EI 2 M0 + V EI 2
= q dx Vu .
dx dx x=L dx dx x=L 0
(1.39)
a(v, ) = L() Vu ,
where
L
d2 v d2
Z
a(v, ) = EI 2 2 dx,
0 dx dx
Z L
d
L() = q dx + M0 V |x=L .
0 dx x=L
27
L 2 L
d2 v
1 dv
Z Z
= EI dx qv dx M0 + V v|x=L . (1.40)
2 0 dx2 0 dx x=L
Now consider the variational formulation for the elasticity problem given
by Eqn. 1.33. The bilinear operator, a(., .) and the linear operator L(.) are
given by
Z
a(u, v) = (v) : C(u) d,
Z Z
v b + (v) : C0 (v) : 0 d.
L(v) = v t d +
t
28
section 1.6. We seek an approximate solution of the form
n
X
u = 0 + cj j ,
j=1
Ac = f , (1.41)
where
Aij = a(j , i ), (1.42)
is a symmetric matrix (since a(., .) is symmetric), and
29
we get Eqn. 1.41. Note however, that when a(, ., ) is not symmetric, we do not
have the existence of the functional I. Hence, deriving the matrix equations
directly from the variational form (V) is preferred.
The approximation functions i should satisfy the following conditions:
1. They should be such that a(i , j ) should be well-defined.
2. The set i , i = 1, . . . , n should be linearly independent.
3. {i } must be complete, e.g., when i are algebraic polynomials, com-
pleteness requires that the set {i } should contain all terms of the
lowest-order order admissible, and upto the highest order desired.
Example:
We shall consider the same example of a beam that we considered in sec-
tion 1.6. Recall that the variational formulation was
a(v, w) = L(w) w Vu ,
where
L
d2 v d2 w
Z
a(v, w) = EI dx,
0 dx2 dx2
Z L
dw
L(w) = qw dx + M0 .
0 dx x=L
Since the specified essential boundary conditions are v = dv/dx = 0 at x = 0,
we can take 0 = 0, and select i such that i (0) = (0) = 0. Choosing
1 = x2 , 2 = x3 and so on, we can write v(x) as
n
X
v(x) = cj j (x); j = xj+1 .
j=1
Ac = f ,
30
On solving for c1 and c2 and substituting in the expression for v(x), we get
5qL2 + 12M0 2
qL 3
v(x) = x x.
24EI 12EI
For n = 3, we get
2 1 2
4 6L 8L c qL + 2M0
1 3
EI 6L 12L2 3 = 1 3
18L c2 qL + 3M0 L .
4
1
8L2 18L3 28.8L4
c3 5
4
qL + 4M0 L 2
which leads to
qx2 2 2 M0 x2
v(x) = (6L 4Lx + x ) + .
24EI 2EI
If the governing equation permits, one can carry out integration by parts,
and transfer the differentiation from u to as in the Rayleigh-Ritz method.
Note, however, that the weighted residual method is more general than the
Rayleigh-Ritz method since no variational form is required. Also note that
the approximating functions in the Galerkin method are of a higher order
than in the Rayleigh-Ritz method. The Rayleigh-Ritz and Galerkin methods
yield identical results for a certain class of problems, e.g., self-adjoint, second-
order ordinary differential equations with homogeneous or non-homogeneous
boundary conditions or the linear elasticity problem. But, in general, they
can yield different results.
31
As an example, consider the ordinary differential equation:
d2 y dy
x2 2
+ 2x 6x = 0,
dx dx
with the boundary conditions y(1) = y(2) = 0. The exact solution of the
above equation is
6
y = + 3x 9.
x
In the Galerkin method, we choose an approximate solution in the form of a
polynomial which satisfies the two essential boundary conditions:
y = c1 1 = c1 (x2 3x + 2).
w = c1 (x2 3x + 2).
One can compute c1 and c2 using the Galerkin method and confirm that the
Rayleigh-Ritz method also gives the same results.
32
1. The method should be convergent.
33
Chapter 2
In this chapter, we discuss the formulation of the finite element method, first
in an abstract setting, and then applied to problems in elasticity and heat
transfer.
2. Assemble the element level matrices to form the global equation. For
example, for a linear statics problem, we get a matrix equation of the
form
Ku = f .
34
freedom to be solved for. Let (1 , 2, . . . , n+m ) be the basis for Vh so that
any v Vh has the unique representation
n+m
X
v= i i , i .
i=1
F (uh ) F (v) v Vh .
or, equivalently,
a(uh , v) = L(v) v Vh . (2.1)
The approximate solution, uh , is of the form
m n
(0 )i i +
X X
uh = i i ,
i=1 i=1
Using the bilinearity of a(., .), we can write the above equation as
n m
a((0 )i , j )i = L(j ) j = 1, n,
X X
a(i , j )i +
i=1 i=1
or in matrix form as
A = b A,
where is an m 1 vector, is an n 1 vector, and
Note that the stiffness matrix A is symmetric and positive definite since
a(., .) is symmetric and V-elliptic:
t A = a(v, v) kvk2 0 .
35
2.3 Finite Element Formulation for Elasticity
In this section, we consider the finite element formulation for the linear elas-
ticity problem based on Eqn. 1.33. Observe that this variational formulation
involves dealing with fourth-order tensors such as the elasticity tensor, C,
whose matrix representation is given by a 3 3 3 3 matrix. Thus, carry-
ing out the assembly of the stiffness matrix and force vector in a computer
implementation based on Eqn. 1.33 would be quite cumbersome. What we
need is an equivalent form of Eqn. 1.33 which would lead to simpler matrix
manipulations in a computer implementation.
Towards this end, using the symmetry of the strain and stress tensors,
we replace them by vectors such that the strain energy density expression
in terms of the stress and strain components remains the same. This leads
to the use of engineering strain components instead of the tensorial compo-
nents. For example, if we consider two-dimensional plane stress/plane strain
problems, then we have
: = xx xx + xy xy + yx yx + yy yy
= xx xx + 2xy xy + yy yy
= tc c ,
c = C(c 0c ) + 0c , (2.2)
36
Z
v t b + [c (v)]t C0c [c (v)]t 0c d
v Vu . (2.3)
Since the above equation is valid for all vectors v, we can write the finite
element formulation as
Find u which satisfies
K u = f , (2.5)
where
Z
K= B t CB d,
37
Z Z Z
f= t
N t d + t
(N b + B t
C0c B t
0c ) d t
B C B d u.
t
38
deriving the finite element formulation, we can directly make use of the vari-
ational statement given by Eqn. 1.37 since the thermal conductivity tensor
is a second-order tensor (as opposed to its counterpart, the elasticity tensor,
which is a fourth-order tensor).
The temperature T written in terms of the interpolation functions is
T = N w + N w,
where w is the vector of the unknown temperatures at the nodes, N are the
shape functions associated with the nodes where the temperature is unknown,
is the vector of the prescribed temperatures at the nodes lying on T , and
w
N are the corresponding shape functions. Noting that the variation, v, of
the temperature vanishes on T , we have
v = N v.
where
Nj
Bij =
xi
Nj
Bij = .
xi
Substituting for these quantities in Eqn. 1.37, and using the arbitrariness of
v, we get
Find w such that
K w = f ,
where
Z Z
t
K= B kB d + hN t N d,
q
Z Z Z
f= t
N Q d + t
N (q + hT ) d t
B kB d w.
q
Note that K is symmetric and positive definite by virtue of the symmetry and
positive-definiteness of k. Also note that the second term in the stiffness
matrix, K, involves an integral over part of the surface of the domain. There
is no corresponding term in the expression for the stiffness matrix in the
elasticity problem. If n is the number of free degrees of freedom and m
the number of prescribed degrees of freedom, then the order of the various
matrices in a three-dimensional problem is as follows: K : n n, w : n 1,
: m 1, N : 1 m.
f : n 1, N : 1 n, B : 3 n, B : 3 m, w
39
t
1 2 P 3 4
L L L
Figure 2.1: Bar subjected to uniaxial loading.
L 2L 3L
Figure 2.2: Example of a piecewise linear function.
Vh H 1 () Vh C 0 (),
40
u
Vh
uh
Figure 2.3: Finite element solution as a projection of the exact solution on
Vh .
x j-1 xj x j+1
Vh H 2 () Vh C 1 (),
where
C 0 () = {v : v is a continuous function on }
1 0 v 0
C () = v C () : C () .
xi
The solution, uh , that we seek is the projection of u on Vh as shown in
Fig. 2.3.
To describe any function in Vh , we introduce the basis functions
(
0 if i 6= j
Nj (xi ) =
1 if i = j
A typical basis function (or shape function) is shown in Fig. 2.4. Since there
are 4 nodes in our finite element model, the basis functions associated with
the nodes, {Ni }4i=1 , form a basis of Vh for this particular problem. Hence,
any function uh Vh can be represented as
4
X
uh = Ni ui ,
i=1
41
Substituting for and , we get the problem statement as
Find uh which satisfies
Z 3L Z 3L Z 3L
dv duh
EA dx = tv dx + bvA dx + P v|x=L v.
0 dx dx 0 0
where
N u v
2 2 2
N = N3 , u = u3 , v = v3 .
N4 u4 v4
Using the fact that v2 , v3 and v4 are arbitrary, we get the usual finite element
equation
K u = f , (2.8)
where
3L 3L 3L
Z
dN2 dN2 dN2 dN3 dN2 dN4
Z Z
dx dx dx
0 dx dx 0 dx dx 0 dx dx
Z 3L 3L 3L
dN3 dN2 dN3 dN3 dN3 dN4
Z Z
K = EA dx dx ,
dx
0 dx dx 0 dx dx 0 dx dx
Z 3L 3L 3L
dN4 dN2 dN4 dN3 dN4 dN4
Z Z
dx dx dx
0 dx dx 0 dx dx 0 dx dx
Z 3L N2 Z 3L N2 P
f= N3 t dx + N3 bA dx + 0 .
0 0
N4 N4 0
Note that
dN2 1 dN3 1 dN4
= 0 x L; = L x 2L; = 0 0 x 2L;
dx L dx L dx
1 1 1
= L x 2L; = 2L x 3L; = 2L x 3L;
L L L
= 0 elsewhere; = 0 elsewhere.
42
Assuming that the tractions and body force, t and b, are constant, the stiff-
ness matrix and force vector are given by
2 1 0
EA
K= 1 2 1
L
0 1 1
(2.9)
2 2 P
tL
bAL
f= 2 + 2 + 0
2 2
1 1 0
Note that the stiffness matrix is banded since any shape function interacts
only with its neighboring shape functions. Thus, for example N2 and N3 are
both nonzero on L x 2L, but N2 and N4 are not both nonzero on any
part of the domain. This leads to the K(1, 2) term being nonzero, but the
K(1, 3) term being zero.
Once Eqn. 2.8 has been solved for u, the strains can be recovered from
the displacement field by using
u
1
u
duh dN1 dN2 dN3 dN4 2
= =
dx dx dx dx dx u3
u4
= B u,
where B is the strain-displacement matrix. We get
1
= (u2 u1 ); 0 x L,
L
1
= (u3 u2 ); L x 2L,
L
1
= (u4 u3 ); 2L x 3L.
L
Note that as a result of assuming a piecewise-linear displacement field the
strain field is constant over each element. Thus, in general, it can be dis-
continuous across element boundaries as shown in Fig. 2.5. If we had as-
sumed a piecewise-quadratic displacement field then we would have obtained
a piecewise-linear (again, not necessarily continuous) strain field. The stress
is simply obtained by using = E. Since the strain is just multiplied by
a constant, the qualitative behavior of the stress field is the same as that of
the strain field.
We would like to caution the reader though, that the procedure of directly
working on the entire domain as just shown is not followed while implement-
ing a finite element program. In an actual implementation, the stiffness and
43
x
0 L 2L 3L
Figure 2.5: The strain field is, in general, discontinuous across element
boundaries.
force matrices are formulated at an element level and then assembled to form
the global stiffness matrix and force vector as shown in the following chap-
ter. The intention behind carrying out the computations at the global level
was to provide insights into issues which get obscured otherwise (e.g., the
banded nature of the stiffness matrix), and also to provide a justification for
the element-level formulation which is followed in practice.
44
Chapter 3
We have seen in the previous chapter that the procedure of formulating the
stiffness and load vectors can be quite cumbersome. It is more convenient to
form the element stiffness and load vectors, and assemble them to form the
global stiffness and load vectors.
N1 N2
1 e 2
45
x1 x2 =1 =+1
u=N1 u1 + N 2 u 2
u2
u1
1 2
N1 () = a + b,
where a and b are constants which are determined from the boundary con-
ditions. We get
1
N1 () = (1 ).
2
Following a similar procedure, we get
1
N2 () = (1 + ).
2
The coordinate x of a point in the element is given by
x = N1 ()x1 + N2 ()x2 ,
u = N1 ()u1 + N2 ()u2.
If the same shape functions are used to interpolate both the displacements
and the geometry as above, the mapping is known as an isoparametric map-
ping. Note that u( = 1) = u1 , u( = +1) = u2 with a linear interpolation
in between as shown in Fig. 3.3.
46
The strain is given by
du du d
= =
dx d dx
1
= (u2 u1 )
x2 x1
1
= (u2 u1 ),
le
where le = (x2 x1 ) is the length of the element. The above equation can
be written as
= B u,
where
1
B= [1 1]
le
u1
u = .
u2
where, now, we have added the contribution of the point loads Pi explicitly
instead of considering it as part of the prescribed tractions t. Writing the
above equation as a summation over the number of elements, we get
XZ X Z Z X
t
[c (v)] c d = v b d + v t d + Pi vi v Vu ,
e e e e e
Note that while computing the surface integral term in the above equation,
we need to consider only the those elements which have a surface on t . The
reason is that the net contribution of the internal tractions at any shared
edge between two elements to the global load vector is zero since the traction
vector on the edge of one element is equal and opposite to the traction vector
on the adjacent element.
For the one-dimensional problem that we are considering, we have v =
N v, (v) = B v, = EB u, d = Adx = Ale d/2 and d = dx = le d/2.
Substituting these various quantities in the above equation, we get
XZ 1 le X Z 1 Z 1
t le le
t t
v B EAB d u = t
v N t d + v N bA d +v t {Ni (j )Pj },
t t
e 1 2 e 1 2 1 2
47
where in the last term j is the position at which load Pj is applied.
From the above equation, the element stiffness matrix is given by
1 1 t
Z
(e)
K = B EABle d.
2 1
The element load vectors due to the traction, body forces and point loads
are
R
Z 1 le 1
(e) 1 N 1 t d
ft = N t tle d = 2 R1 ,
2 1 le 1
N t d
2 1 2
R
Z 1 le 1
(e) 1 N 1 bA d
fb = N t Able d = 2 R1 ,
2 1 le 1
N bA d
2 1 2
P
(e) i N1 (i )Pi
f P = P .
i N2 (i )Pi
48
3.2 Assembly of the Element Stiffness and
Load Vectors into the Global Stiffness and
Load Vectors
Consider the contribution of the second element to the term v t K u in Eqn. 2.4.
We have
2 2 1 1 u2
h iE A
(v t K u)2 = v2 v3
l2 1 1 u3
0 0 0 0 u1
h i 0 E 2 A2 E 2 A2 0
u2
l l
= v1 v2 v3 v4 2 2 .
0 E2l A2 E2l A2 0 u3
2 2
0 0 0 0 u4
Thus, we see that the elements of K (2) occupy the second and third rows
and columns of the global stiffness matrix K. Since two and three are the
global degrees of freedom associated with element 2, we conclude that the
element stiffness matrix should be assembled based on the degrees of freedom
associated with the element. Overlapping elements are added. Similar to the
process for the stiffness matrix, we assemble the global load vector using the
individual element load vectors. For the model 1-d problem considered in
Section 2.5, we get
1 1 0 0
1 2 1 0
EA
K=
L 0 1 2 1
0 0 1 1
(3.1)
1 1 R
2 bAL 2 P
tL
f=
+ 2 + ,
2 2 2 0
1 1 0
49
1. The dimension of the stiffness matrix is n n where n is the number
of free degrees of freedom.
2. K is symmetric.
nbw = max(4 1, 4 3, 3 2) + 1 = 4,
50
the external force applied and the reaction corresponding to this degree of
freedom are F1 and R1 , respectively. Then we have
K K12 . . . K1n a F + R1
11 1 1
K21 K22 . . . K2n u2 F2
= .
. . . . . . . . . . . . . . . . . . . . . .
Kn1 Kn2 . . . Knn un Fn
or, alternatively,
and so on. In matrix form, all the equations excluding the first can be written
as
K22 K23 . . . K2n u2 F2 K21 a1
K32 K33 . . . K3n u3 F3 K31 a1
=
.
. . . . . . . . . . . . . . . . . . . . . . . ...
Kn2 Kn3 . . . Knn un Fn Kn1 a1
(n1)(n1) (n1)1 (n1)1
Thus, the reduced K matrix is obtained by eliminating the row and column
corresponding to the specified or support degree of freedom. Note that the
force matrix also gets modified when the prescribed quantity is non-zero.
Since the rigid-body modes have been eliminated, the reduced K matrix is
no longer singular. Hence we can solve for the unknown vector u.
In order to compute the strains the element-level quantities are extracted
from the global u vector. We then use = B u(e) . The stress is then obtained
using the appropriate constitutive relation, e.g., in our model problem =
EB u(e) . In order to compute the reactions, we use Eqn. 3.3 since all the
quantities on the right-hand side of that equation are known. Thus, if we
need to compute the reactions, then we need to store the rows of K which
are eliminated for the purposes of getting the reduced stiffness matrix.
The above procedure can be easily generalized for the case when there
are r prescribed degrees of freedom, say up1 = a1 , up2 = a2 , . . ., upr = ar :
51
C
a1
Store the p1 , p2 , . . ., pr row of the global stiffness matrix and the load
vector f .
Delete the p1 , p2 , . . ., pr th row and column from the K matrix. The
resulting stiffness matrix size is (n r) (n r). Delete the p1 , p2 , . . .,
pr th row from the load vector. Modify the load components as
X
fi fi Ki,pj aj .
j
Solve K u = f .
Extract the element displacement fields u(e) from u, and determine the
strains and stresses.
Evaluate the reactions at each support degree of freedom:
X
Rp1 = Kp1 1 u1 + Kp1 2 u2 + . . . + Kp1 n un fp1 = Kp1 i ui fp1 ,
i
X
Rp2 = Kp2 1 u1 + Kp2 2 u2 + . . . + Kp2 n un fp2 = Kp2 i ui fp2 ,
i
One good debugging check is to see that the (vectorial) sum of the
reactions is equal to the load applied on the structure.
In the model problem considered in Section 2.5, the reaction at the wall is
given by
1 1 1
R = EAu2 ALb Lt.
L 2 2
53
3.3.3 Multi-point constraints
Multi-point constraints are constraints involving several displacement degrees
of freedom, as, for example, occurs when a node is constrained to move along
an inclined surface. We consider only linear multi-point constraints in what
follows. Let the constraints be of the form
EC u = Ea, (3.5)
K u EC T = f . (3.6)
3.3.4 Example
To illustrate the elimination and penalty approaches, consider the example
shown in Fig. 3.5. The Youngs modulus of the two bars are given by E1 =
70 109 N/m2 and E2 = 200 109 N/m2 . The cross-sectional areas are
A1 = 2400 mm2 and A2 = 600 mm2 . The applied load is P = 200 103 N.
The element stiffness matrices are (with units N/mm)
3 1 1
70 10 2400
K (1) =
300 1 1
54
300 mm 400 mm
1 2 P 3
E1 E2
3 1 1
200 10 600
K (2) =
400 1 1
Using the elimination approach, we the 1st and 3rd rows and columns from
K. We get
106 (0.86)u2 = 200 103 ,
which yields u2 = 0.23255 mm. The element stresses are
E1 h i u1
1 = E1 B 1 u(1) = 1 1 = 54.77 MPa,
(x2 x1 ) u2
E1 h i u2
2 = E2 B 2 u(2) = 1 1 = 116.29 MPa.
(x2 x1 ) u3
Note that the sum of the reactions is equal to the applied load.
The results for (u1 , u2 , u3, R1 , R2 ) could directly have been obtained using
Eqns. (3.7), with (R1 , R2 ) playing the role of the Lagrange multipliers.
Now let us carry out the same computations using the penalty approach.
Using the suggested guidelines, we have C = 0.8 1010 . The stiffness matrix
is modified by adding C to the appropriate diagonal terms. The matrix
equations are given by
8600.56 0.56 0 u 0
1
106 0.56 0.86 0.3 u2 = 200 103 .
0 0.3 8600.3 u3 0
55
N1 N2 N3
=1 e =1 =1 e =1 =1 =1
e
Figure 3.6: Quadratic shape functions: Global and element level pictures
56
expression for N1 . Since N1 is zero at = 0 and = 1, we assume N1 =
c(1 ). The constant c is determined by using the condition N1 |=1 = 1.
A similar procedure is followed for determining N2 and N3 .
Note that the shape functions formulated above give C 0 continuity for the
function being interpolated, but not, in general C 1 continuity. Such shape
functions are called Lagrange shape functions1 . In general, a function of
degree (n 1) and defined by n values i at corresponding i has the form
= N1 1 + N2 2 + . . . + Nn n ,
where
(2 )(3 )(4 ) . . . (n )
N1 = ,
(2 1 )(3 1 )(4 1 ) . . . (n 1 )
(1 )(3 )(4 ) . . . (n )
N2 = ,
(1 2 )(3 2 )(4 2 ) . . . (n 2 )
...
(1 )(2 )(3 ) . . . (n1 )
Nn = .
(1 n )(2 n )(3 n ) . . . (n1 n )
For n = 3, we can easily verify that we get the expressions in Eqn. 3.8.
For an isoparametric element, we have
u = N1 u1 + N2 u2 + N3 u3 ,
x = N1 x1 + N2 x2 + N3 x3 ,
u = N u
x = N x.
57
u
x x
u
2 h i 1
= dN1 dN2 dN3
u2
(x3 x1 ) d d d
u3
= B u(e) ,
where
2 h 12 1+2
i
B= 2 2 .
le 2
Note that the size of the element stiffness matrix is ndof ndof , where ndof is
the number of element degrees of freedom.
The element force vector due to a constant body force b is
Z x2
(e)
f = N t bA dx
x1
Z 1
le
= N t bA d
1 2
58
5
cm
0
3
10
x 2
cm
1
50
Figure 3.8: Analysis of a rotating bar
1
bAle
Z
= N t d
2
1
1
bAle
= 4 .
6
1
Similarly, the element-force vector due to a constant force per unit length t
is
1
(e) l t
e
ft = 4 .
6
1
3.5 Examples
Example 1:
A rod of length 150 cm rotates about a point with a constant angular velocity
as shown in Fig. 3.8. The density, Youngs modulus, cross sectional area, and
angular velocity are given by = 7500 kg/m3 , E = 60 GPa, A = 4104 m2 ,
and = 20c /s. Assume u = u(x)ex , and the loading to be only due to the
centrifugal body force (given by b = x 2 ). Formulate the element stiffness
matrix, and the consistent load vector for a quadratic (3-node) element. Then
using the two-element (5-node) model shown in the figure, determine the
displacements and the stress distribution in the rod using two quadratic
elements. Plot the stress distribution as a function of x. What do you
observe about the stress distribution at node 3?
Solution:
59
The body force is given by
x
h i 1
b = x 2 = 7500(400) N1 N2 N3 x2 .
x3
The global force vector obtained after assembling the two load vectors is
0
100
f = 150 .
800
300
60
As a check, we verify that the total load obtained by summing the entries
in the load vector (1350 N) is the same as obtained from the continuum
problem, i.e., Z 1.5
1.2 103 x dx = 1350 N.
0
The element stiffness matrix is given by Eqn. 3.9. For the two elements,
we have
14 16 2 7 8 1
(1) 7 (2) 7
K = 0.8 10 16 32 16 ; K = 0.8 10 8 16 8 .
2 16 14 1 8 7
61
3437500 3375000
3281250
3250000
3250000 3000000
3062500 1750000
1875000
250000
x x
(a) (b)
Figure 3.9: Stress field in rotating bar: (a) Finite element & (b) Exact
solutions
We shall use this load vector to solve the statically indeterminate problem
shown in Fig. 3.10. The material and geometric properties are E1 = 70 GPa,
A1 = 900 mm2 , 1 = 23 106 /o C, E2 = 200 GPa, A2 = 1200 mm2 , 2 =
11.7106 /o C. The point load and change in temperature are P = 3105 N
and T = 40o C.
The element stiffness matrices are
1 1 3 1 1
E1 A1 = 70 10 900
K (1) = ,
l1 1 1 200 1 1
62
T = 40 o C
E1, A1, 1 P E 2 , A2 , 2
200 mm 300 mm
E2 A2 1 1 3 1 1
K (2) = = 200 10 1200 .
l2 1 1 300 1 1
The global load vector due to the thermal loads (obtained using Eqn. 3.10)
and the point load P is given by
57.96 + R1
3
f = 10 245.64 .
112.32 + R2
which on solving yields u2 = 0.22 mm. The global displacement vector is,
thus, [0 0.22 0]t mm. The stress in the bars is
Note from the above example that statically indeterminate problems can be
handled in a routine way by the finite element method.
63
u2
6 10
u4
5 9
(x 2, y2 )
u3
8 u1 le
2 4 12
u2
1 3 7
11
(x 1, y1 )
u1
Figure 3.11: Global and local degrees of freedom for a truss member.
u1 = u1l + u2 m,
u2 = u3l + u4 m,
where l = cos and m = sin are the direction cosines which are given by
x2 x1 y2 y1 p
l= ; m= ; le = (x1 x2 )2 + (y1 y2 )2 .
le le
In matrix form, we have
u
1
u u
1 = Q 2 , (3.11)
u2 u3
u4
where
l m 0 0
Q= .
0 0 l m
In the local coordinate system, the element stiffness is given by
Ee Ae 1 1
K = .
le 1 1
The element stiffness matrix in global coordinates is derived using the fact
that by virtue of the tensorial nature of the quantities involved, the virtual
work contribution is the same irrespective of the coordinate system that we
work in. Thus,
(v )t K u = v t (Qt K Q)u,
leading to
K (e) = Qt K Q
64
l2 lm l2 lm
2 2
lm m lm m
Ee Ae
= .
le l 2
lm l 2
lm
2 2
lm m lm m
One gets the same result by equating the strain energy in the local and global
coordinate systems.
The stress in a member is given by
Ee h i u1
= 1 1 Ee e T
le u2
u
1
u2
i
Ee h
= l m l m
Ee e T.
le u3
u4
The thermal load vector is obtained by equating the virtual work expres-
sions in the local and global coordinate systems. We get
Z 1
t t t le
(v ) f th = (v) Q B t Ee e T Ae d,
1 2
65
3.7 One-Dimensional Heat Transfer Problems
Since the procedure for formulating the stiffness and load vectors is exactly
analogous to the elasticity problem, we shall illustrate it directly by means
of an example.
Problem: Find the temperature distribution in a bar of length L subject
to the following boundary conditions:
1. T (0) = T1 ; T (L) = T2 .
2. k dT = q; k dT
dx x=0 dx
+ h(T T ) x=L = 0.
3. T (0) = T1 ; k dT
dx x=L
= q.
Solution: If we assume linear elements, then
T = N1 w1 + N2 w2 ,
where N1 = (1 )/2 and N2 = (1 + )/2. Thus,
dT h i w 1 w
= dN 1 dN2 = B 1
dx dx dx
w2 w2
1. For the boundary conditions given, we see that = T . Thus,
Z 1
l kA 1 1
K (e) = kB t B A d = ,
1 2 l 1 1
where l denotes the length of the element. If we divide the domain into
3 elements of equal length (l = L/3), then the finite element equations
prior to incorporating the boundary conditions are
1 1 0 0 T1 1 q
1
1 2 1 0 w2 QAl 2 0
kA
= + A ,
l 0 1 2 1 w3
2 2
0
0 0 1 1 T2 1 q4
where q1 and q4 are the unknown heat fluxes at the two ends. The
equations for w2 and w3 can be written as
kA 2 1 w2 1 kA T1
= QAl + ,
l 1 2 w3 1 l T2
66
Solving for w2 and w3 , we get
QL2 1
w2 = + (2T1 + T2 ),
9k 3
QL2 1
w3 = + (T1 + 2T2 ),
9k 3
which on substituting in the equations for q1 and q4 yields
1 k
q1 = QL + (T1 T2 ),
2 3
1 k
q4 = QL + (T2 T1 ).
2 3
The exact solution is
QL2 x x 2
1
T (x) = + (T2 T1 )x + T1 ,
2k L L L
dT QL 2x k
q(x) = k = 1 (T2 T1 ).
dx 2 L L
Note that the solution for the temperature field is exact at the nodes.
However, since we have assumed a linear interpolation for the temper-
ature, while the real variation is quadratic, there is an error in the
finite element solution at points other than the nodes. Since the exact
solution is quadratic, we can get the exact solution simply using one
quadratic element. We now have
7 8 1 T 1 q
kA 1 AQl 1
8 16 8 w2 = 4 + 0 ,
3l 6
1 8 7 T2 1 q3
where l = L. The solution is
1 QL2 1
w2 = + (T1 + T2 ).
8 k 2
The heat flux at the ends is
1 k
q1 = QL + (T1 T2 ),
2 L
1 k
q3 = QL + (T2 T1 ).
2 L
2. The two ends of the bar constitute q . The finite element equations
are
1 1 0 0 w 1 Aq
1
1 2 1 0 w2 AQl 2 0
kA
= +
l
0 1 2 1 w3
2 2 0
hl
0 0 1 1 + k w4 1 AhT
67
On solving, we get
QL + q L(QL + 2q)
w1 = + + T ,
h 2k
QL + q 2L(2QL + 3q)
w2 = + + T ,
h 9k
QL + q L(5QL + 6q)
w3 = + + T ,
h 18k
QL + q
w4 = + T .
h
The exact solution is
QL2 x2
2k qL k x
T (x) = 1+ + 1+ + T .
2k hL L2 k hL L
Note again that the temperature solution is exact at the nodes.
3. The right end of the bar constitutes q . The finite element equations
are
1 1 0 0 T 1 Aq
1 1
1 2 1 0 w2 AQl 2 0
kA = + ,
l 0 1 2 1 w3
2 2 0
0 0 1 1 w4 1 Aq
which on incorporating the boundary conditions becomes
2 1 0 w2 2 0 T1
kA
AQl kA
1 2 1 w3 = 2 + 0 + 0 .
l 2 l
0 1 1 w4 1 Aq 0
On solving, we get
L(5QL + 6q)
w2 = + T1 ,
18k
2L(2QL + 3q)
w3 = + T1 ,
9k
L(QL + 2q)
w4 = + T1 .
2k
The flux at the left end is
k 1
q1 = (T1 w2 ) Ql = (QL + q).
l 2
The exact solution is
L2 Q 2x x2
qx
T (x) = 2 + 2 + T1 ,
k 2 L L L
dT h x i
q(x) = k = QL 1 + q .
dx L
Note that the temperature solution is exact at the nodes.
68
3.8 Beams and Frames: An Introduction to
C 1 elements
We shall consider only the classical or slender beam theory here, and not the
Timoshenko beam theory which includes the effect of shear deformation. We
restrict our attention to beams with symmetric cross-sections.
Assuming that the x axis is directed along the length, and the y-axis is
oriented downward with y = 0 at the neutral plane, we have
My
xx = = Exx .
I
The remaining stress components are assumed to be zero. The governing
differential equation for the displacement is
d2 v
EI = M(x),
dx2
where M(x) is the bending moment along the beam.
The expression for the potential energy is
1
Z Z Z X
t
= c d t u d b u d Pi u i
2 t
1 L M2
Z Z Z X X
2
= 2
y dA dx qv dx p i vi Mi vi
2 0 EI A
Z L 2 2 Z L
1 d v X X
= EI dx qv dx P i vi Mi vi .
2 0 dx2 0
The variational formulation can also be obtained from the governing equa-
tions
dM dV d2 v
= V ; = q; EI 2 = M,
dx dx dx
where V is the shear force. The above equations lead to the governing equa-
tion
d2 d2 v
EI 2 q = 0.
dx2 dx
Multiplying this governing equation by a test function , and integrating
by parts leads us to the variational formulation as has already been demon-
strated in Section 1.6.
In order that d2 v/dx2 is well-defined, we need v C 1 , i.e, both v and
v must be continuous. Now we formulate Hermite shape functions which
69
u1 u3
=1 u =+1 u
2
=0 4
Slope=0
N1
1 N2
Slope=0 Slope=0
Slope=1
Slope=0
N3 Slope=0 Slope=1
1
Slope=0 N4
v() = N1 v1 + N2 v1 + N3 v2 + N4 v2 ,
where
dv dv
v1 = ; v2 = .
d =1 d =+1
In order that v(1) = v1 , v (1) = v1 , v(+1) = v2 and v (+1) = v2 , the
shape functions should be as shown in Fig. 3.13. We assume each of the Ni s
to be of the form
Ni () = ai + bi + ci 2 + di 3 ,
and find the constants ai -di based on the boundary conditions that each shape
function has to satisfy. For example, N1 (1) = 1, N1 (1) = 0, N1 (+1) = 0,
and N1 (+1) = 0. We get
1 1
N1 = (1 )2(2 + ) = (2 3 + 3 ),
4 4
70
1 1
N2 = (1 )2(1 + ) = (1 2 + 3 ),
4 4
1 1
N3 = (1 + )2 (2 ) = (2 + 3 3 ),
4 4
1 1
N4 = (1 + )2 ( 1) = (1 + 2 + 3 ).
4 4
We assume that the coordinate interpolation is still given by
1 1
x = (1 )x1 + (1 + )x2 ,
2 2
leading to dx = le d/2. Using the chain rule
dv dv dx le dv
= = .
d dx d 2 dx
le le
v() = N1 u1 + N2 u2 + N3 u3 + N4 u4.
2 2
In matrix form, the above equation can be written as
v = N u(e) , (3.12)
where
h i
l l
N N1 2 N2 N3 2 N4 ,
e e
h it
u(e) u1 u2 u3 u4 .
d2 v 4 d2 v
= .
dx2 le2 d 2
Using the same interpolation for the variations and substituting in the vari-
ational formulation, we get
Z 1
16EI d2 N t d2 N le
Z 1
t t t le
X X
4 2 d 2 2
d u = N q d+ P i (xi )+ Mi (xi ).
1 l e d 1 2
where
1
le
Z
(e)
K = B t (EI)B d
1 2
71
1
le
Z
(e)
f = N tq d + {Ni (j )Pj } + {Ni (j )Mj },
1 2
with B given by
4 d2 N 4 h3 le
i
B= 2 = 2 2 (3 1) 23 le
(3 + 1) .
le d 2 le 4 4
Assuming that q is uniform over the element, and that there are no point
loads and moments, the element load vector is
qle
2
qle2
12
f (e)
= .
qle
2
2
ql12e
The element level matrices are assembled into the global ones as usual,
and we solve the system of equations K u = f . The moment and shear force
are found as follows:
d2 v
M = EI
dx2
d2 N
= EI 2 u(e)
dx
4EI d2 N (e)
= 2 u
le d 2
EI
= 2 [6u1 + (3 1)le u2 6u3 + (3 + 1)le u4 ] ,
le
dM
V =
dx
2 dM
=
le d
6EI
= 3 [2u1 + le u2 2u3 + le u4 ] .
le
72
12 kN/m
1m 1m
73
which on solving yields u4 = 2.679 104 radians and u6 = 4.464
104 radians. The reactions are given by
The positive sign on R1 indicates that the reaction at the wall is directed
downwards, while the negative signs on R3 and R5 indicate that the reactions
at the roller supports are directed upwards. The positive sign on R2 indicates
that the moment is clockwise at the wall.
The deflection at the midpoint of the second element is
v = N u(2)
=0
0
iu4
h
= 0.5 0.5(0.25) 0.5 0.5(0.25)
0
u6
= 0.0893 mm.
V = 6EI u4 = R1 ,
as expected.
u = Qu,
74
u5
u5 u4
u4
u 6 (u6 )
u2
u2
u1
u1
u 3 (u3)
where
l m 0 0 0 0
m l 0 0 0 0
0 0 1 0 0 0
Q= .
0 0 0 l m 0
0 0 0 m l 0
0 0 0 0 0 1
Similar to the truss element, the element stiffness matrix in global coordinates
is given by
K (e) = Qt K Q,
where
EA
l
0 0 EA
l
0 0
12EI 6EI
0 0 12EI 6EI
l3 l2 l3 l2
6EI 4EI
0 0 6EI 2EI
l2 l l2 l
K =
.
EA EA
l 0 0 0 0
l
0 12EI 6EI 0 12EI
6EI
l3 l2 l3 l2
6EI 2EI
0 l2 l
0 6EI
l2
4EI
l
f (e) = Qt f ,
75
3.10 A Note on Equation Solving
We have seen that our finite element equations are a system of algebraic
equations of the form
K u = f .
There are two types of methods available for solving these equations, namely
(i) Direct methods which are based on some variant of Gauss elimination,
(ii) Iterative methods.
In the direct methods, K is decomposed into a lower triangular matrix,
L with unit diagonals, and an upper triangular matrix U , i.e.,
K = LU .
Lx = f and U u = x.
Thus,
x1 = f1
i1
X (3.13)
xi = fi Lij xj i = 2, . . . , n.
j=1
xn
un = ,
Unn
(3.14)
" n
#
1 X
ui = xi Uij xj i = n 1, n 2, . . . , 1.
Uii j=i+1
Eqn. 3.13 is called forward elimination, while Eqn. 3.14 is called back sub-
stitution. The actual decomposition K = LU is carried out using some
variation of Gauss elimination.
Once the L-U decomposition is computed, several solutions for different
load vectors can be computed using Eqns. 3.13 and 3.14. This process is
called resolution since it is not necessary to recompute L and U . Note
that the L-U decomposition is very costly compared to the resolution (o (n3 )
versus o (n2 )).
Several strategies exist for taking advantage of the banded and symmetric
form of the stiffness matrix. Even further storage reductions are possible by
using a profile of the K matrix. We then need to use a profile equation
solving program. See [4] for details of such solvers and of iterative techniques
such as the Gauss-Seidel method.
76
Chapter 4
Modeling of Two-Dimensional
Problems
u6
4 5 6
8 7
(x 3, y3 ) u5
8 7 4 1
2
3 1 u2 u4
u1 u3
(x 1, y1 ) (x 2, y2 )
77
=0
N
=0
=0
=0
=0
1 =0
3
1
1
=1 =1
2
78
=0
3
A2 A1
=1
A3 2
1 =0
=1
Figure 4.3: Representation of shape functions using area coordinates
have
u = N1 u1 + N2 u3 + N3 u5 ,
(4.4)
v = N1 u2 + N2 u4 + N3 u6 ,
or, alternatively,
u = N u,
where
N1 0 N2 0 N3 0
N =
0 N1 0 N2 0 N3
h it
u = u1 u2 u3 u4 u5 u6 .
x = N1 x1 + N2 x2 + N3 x3 ,
(4.5)
y = N1 y1 + N2 y2 + N3 y3 .
u = (u1 u5 ) + (u3 u5 ) + u5 ,
79
v = (u2 u6 ) + (u4 u6 ) + u6 ,
x = (x1 x3 ) + (x2 x3 ) + x3 ,
y = (y1 y3 ) + (y2 y3 ) + y3 ,
We now turn to the evaluation of the B matrix. Using the chain rule, we
have
u u x u y
= + ,
x y
u u x u y
= + ,
x y
where
1 y23 y13
J 1 =
det J x23 x13
80
4.8 = 2N1 + 3.5N2 + 7N3 ,
1 = N1 + N2 + N3 .
Now using the fact that y13 = y31 , y12 = y21 and so on, we get
y23 0 y31 0 y12 0
1
B= 0 x32 0 x13 0 x21
det J
x32 y23 x13 y31 x21 y12
where the last step follows from the fact that B and C are constant over the
element. Using Eqn. 3.2, the half-bandwidth is given by
nbw = 2 max me + 1 .
1ene
81
In the above equation me is given by
me = max{|i1 i2 | , |i2 i3 | , |i1 i3 |}
where i1 , i2 and i3 are the node numbers of an element.
The consistent force vector due to body forces is given by
Z
fb = N t bte dA.
Ae
82
u4
u3
=1 2
u2
l
u6
1 u1
=0 3 u5
=1
=0
Ae
= .
3
The contribution to the load vector due to the tractions acting on an edge
of the element is given by
Nt
1 x
N1 ty
N t
Z Z
t
2 x
f t = N tte dA = te ds.
N2 ty
N3 tx
N3 ty
As an example consider the case shown in Fig. 4.4. On edge 1-2, we have
x = x1 + x2 ,
y = y1 + y2 ,
leading to
dx = dx1 + dx2 ,
dy = dy1 + dy2 .
83
4
10000 N
3 8
2 4
7
2
30 mm
1
2 6
1 3 5
1
60 mm
Figure 4.5: Example problem
Therefore R
1
t N l d
x R0 1 e
1
ty 0 N1 le d
R
tx 01 N2 le d
f t = te
R1
.
ty 0 N2 le d
R
1
tx 0 N3 le d
R
1
ty 0 N3 le d
Using the fact that one edge 1-2, N1 = , N2 = = 1 and N3 = 0, we get
t
x
ty
t
te l
x .
ft =
2 ty
0
0
11 3 2
84
2 3 4 2.
The coordinates of the nodes 1-4 are (0.0, 0.0), (0.0, 0.03), (0.06, 0.0) and
(0.06, 0.03). For element 1, we have x13 = 0, y23 = 0.03, x23 = 0.06,
y13 = 0.03, y12 = 0, x21 = 0.06, det J = 0.06(0.03) = 0.0018. Thus,
0.03 0 0.03 0 0 0
(1) 1
B = 0 0.06 0 0 0 0.06 .
0.0018
0.06 0.03 0 0.03 0.06 0
For element 2, we have x13 = 0.06, y23 = 0, x23 = 0.06, y13 = 0.03,
y12 = 0.02, x21 = 0, det J = 0.06(0.03) = 0.0018, yielding
0 0 0.03 0 0.03 0
(2) 1
B = 0 0.06 0 0.06 0 0 .
0.0018
0.06 0 0.06 0.03 0 0.03
85
The global degrees of freedom associated with the two elements are (1, 2, 5, 6, 3, 4)
and (5, 6, 7, 8, 3, 4). Assembling the element level matrices into the global
stiffness matrix and incorporating the boundary conditions, we get
2.1 0 1.2 0.6 u5 0
0 3.9 0.6 3.6 u6 0
7
9 2.43 10 =
.
1.2 0.6 2.1 1.2 u7 0
0.6 3.6 1.2 3.9 u8 10000
N1 = c(1 )(1 ),
86
u6 (1, 1) (1, 1)
u8 u5
u7
u2 u4
u1 (1, 1) (1, 1)
u3
Figure 4.6: The bilinear element: Mapping the actual element into a master
element
which ensures that it is zero on the edges 2-3 and 3-4. The constant c is
determined by using the fact that N1 is one at (, ) = (1, 1). A similar
procedure is followed for finding N2 , N3 and N4 . We get
1
N1 = (1 )(1 ),
4
1
N2 = (1 + )(1 ),
4
1
N3 = (1 + )(1 + ),
4
1
N4 = (1 )(1 + ).
4
A general way of representing the above functions is
1
Ni = (1 + i )(1 + i ),
4
where (i , i ) are the coordinates of node i.
The displacement field can be written as
u = N1 u1 + N2 u3 + N3 u5 + N4 u7 ,
v = N1 u2 + N2 u4 + N3 u6 + N4 u8 ,
or alternatively in matrix form as u = N u, where
N1 0 N2 0 N3 0 N4 0
N = .
0 N1 0 N2 0 N3 0 N4
87
where
x y
J11 J12
J =
x y
J21 J22
x1 y1
1 (1 ) (1 ) (1 + ) (1 + ) x2 y2
= .
4 (1 ) (1 + ) (1 + ) (1 ) x3 y3
x4 y4
Thus,
f f
11 12
x = ,
f f
y
21 22
where
J22 J12 J21 J11
11 = ; 12 = ; 21 = ; 22 = ,
|J | |J | |J| |J |
where |J | det J .
The strain-displacement matrix is found as follows:
We have (using the comma notation for denoting partial derivatives)
u,x u
xx 1 0 0 0 ,x
u,y u
c = yy = 0 0 0 1
= R1 ,y . (4.6)
v,x
v,x
xy 0 1 1 0
v,y v,y
Next, we have
u,x 11 12 0 0 u, u,
u,y 21 22 0 0 u, u
= = R2 , . (4.7)
v,x 0 0 11 12 v, v,
v,y 0 0 21 22 v, v,
88
Finally, we have
u
1
u
2
u N 0 N2, 0 N3, 0 N4, 0 u3
, 1,
u, N1, 0 N2, 0 N3, 0 N4, 0 u4
=
v, 0 N1, 0 N2, 0 N3, 0 N4, u5
v, 0 N1, 0 N2, 0 N3, 0 N4, u6
u7
u8
(1 ) 0 (1 ) 0 (1 + ) 0 (1 + ) 0
(1 ) 0 (1 + ) 0 (1 + ) 0 (1 ) 0
1
= u
4
0 (1 ) 0 (1 ) 0 (1 + ) 0 (1 + )
0 (1 ) 0 (1 + ) 0 (1 + ) 0 (1 )
= R3 u. (4.8)
c = R1 R2 R3 u,
thus yielding
B = R1 R2 R 3 .
The stiffness matrix is given by
Z Z 1 Z 1
t
K= B CB d = B t CBt |J | d d.
1 1
The body force term is most easily computed using Gaussian quadrature. As
an example, consider the computation of the consistent load vector due to
a uniform traction on the edge 2-3 in Fig. 4.6. Since x = N2 x2 + N3 x3 and
y = N2 y2 + N3 y3 with N2 = (1 )/2 and N3 = (1 + )/2 along this edge,
we have dx = (x3 x2 ) d/2 and dy = (y3 y2 ) d/2 leading to
p tl23
d = t ds = t dx2 + dy 2 = d,
2
89
=0
=1 3
=1/2
5 2 6
=1/2 5 =1
3
1 4
6 =0 2
4
=1 =1/2 =0
1
90
3 4 7 3
7 1++=0
1+=0
4 6
8 6
8 1=0
2 1+=0
5 1 5 2
1
and then determining k. The final expressions for the shape functions are
1 1
N1 = (1 )(1 )(1 + + ), N5 = (1 2 )(1 ),
4 2
1 1
N2 = (1 + )(1 )(1 + ), N6 = (1 + )(1 2 ),
4 2
1 1
N3 = (1 + )(1 + )(1 ), N7 = (1 2 )(1 + ),
4 2
1 1
N4 = (1 )(1 + )(1 + ), N8 = (1 )(1 2 ).
4 2
The stiffness matrix is given by
Z 1Z 1
(e)
K = B t CBt |J | d d,
1 1
91
4 7 3
3
7
4 6 8 9 6
9
8
2 1 5 2
5
1
93
4.4.1 One-dimensional problems
Consider an integral of the form
Z 1
I= f () d.
1
We approximate it as
I = w1 f (1 ) + w2 f (2 ) + . . . + wn f (n ), (4.9)
where w1 , w2 ,...,wn are the weights, and 1 , 2 ,..., n are the sampling points
or Gauss points. We select the Gauss points and weights such that Eqn. 4.9
provides an exact answer for polynomials of as large a degree as possible. The
idea is that if n-point formula is exact for polynomials of as high a degree as
possible, then the formula will work well even if f () is not a polynomial.
One-point formula
We approximate the integral just using the first term, i.e.,
Z 1
f () d w1 f (1 ).
1
The two parameters w1 and 1 are found such that the integral is exact when
f () is of the form a0 + a1 . Hence, we get
Z 1
(a0 + a1 ) d = w1 (a0 + a1 1 ),
1
or, alternatively,
2a0 = w1 a0 + w1 1 a1 .
Comparing the coefficients of a0 and a1 , we get w1 = 2 and 1 = 0. Thus,
the one-point approximation formula for any arbitrary f () is
Z 1
f () d 2f (0).
1
Two-point formula:
Now we have the approximation
Z 1
f () d = w1 f (1 ) + w2 f (2 ).
1
94
Comparing the coefficients of a0 , a1 , a2 and a3 , we get
w1 + w2 = 2,
w1 1 + w2 2 = 0,
2
w1 12 + w2 22 = ,
3
w1 13 + w2 23 = 0.
Solving the above equations, we get w1 = w2 = 1, and 1 = 2 = 1/ 3.
Thus, for any arbitrary function f (), we have
Z 1
1 1
f () d f +f .
1 3 3
In general, an n-point Gauss-quadrature formula will provide an exact
answer if f () is a polynomial of order (2n 1) or less.
Note that the Gauss points are located symmetrically with respect to the
origin,and that symmetrically placed points have the same weight.
Example:
Consider the approximate evaluation of
Z 1
x 2 1
I= 3e + x + dx = 8.8165.
1 x+2
95
( 13 , 13 ) ( 13 , 13 )
4 3
1 2
( 13 , 13 ) ( 13 , 13 )
is approximated as
" n #
Z 1 X
I= wi f (i , ) d
1 i=1
n
" n #
X X
= wj wi f (i , j )
j=1 i=1
n
XX n
= wi wj f (i , j ).
i=1 j=1
x = N1 x1 + N2 x2 + N3 x2 + N4 x1
1 1
= (1 )x1 + (1 + )x2
2 2
1 1
y = (1 )y1 + (1 + )y2 .
2 2
96
The Jacobian is
x y 1
(x x1 ) 0
J=
= 2 2 = constant.
x y 1
0 (y
2 2
y1 )
B t CB |J| = f ( 2 , 2 , ).
Since the highest order of the polynomial in one direction is 2, the order of
Gaussian integration which is needed is 2 2 (2n 1 2 = n = 2). In
a similar way, the required integration order for evaluating the mass matrix
or the load vector can be assessed.
If the element is distorted, i.e., it is not rectangular or a parralelogram,
then the Jacobian matrix is not a constant over the element. In such a case,
the integrand in the stiffness matrix is a ratio of polynomials which in general
is not a polynomial. Full numerical integration is the order of integration
that gives exact matrices (i.e., same as analytical integration) when the ele-
ments are undistorted. Using this integration order for a distorted element
will not yield the exactly integrated element matrices. However, the numer-
ical integration errors are small if the geometric distortions are small. If the
element is highly distorted (a situation which should be avoided as far as
possible), one should use a higher order of quadrature than the one used in
full integration. The recommended full gauss numerical integration orders
for the 4, 8, 9 and 16 node isoparametric quadrilateral elements are 2 2,
3 3, 3 3 and 4 4, respectively.
If an integration order lower than the full order is used it is called re-
duced integration. Using reduced integration is, in general, unreliable. It
can result in an instability known variously as a mechanism, kinematic mode,
hourglass mode or zero-energy mode. For example, the nodal displacement
in the setup shown in Fig. 4.11 where 2 2 integration is used to compute
the stiffness matrix are very large due to the zero-energy mode. An eigen-
value analysis of the element stiffness matrices for a plane [solid] element
should yield three [six] zero eigenvalues corresponding to the three [six] rigid
body modes (2 translations and a rotation [3 translations and 3 rotations]).
However, in the problem shown above, the element stiffness computed using
a 2 2 integration rule has one spurious zero-energy mode, i.e., four zero
eigenvalues, due to the use of reduced integration.
97
P
where wi , i and |J |i are the values of the weight, function and determinant
of the Jacobian at the sampling point. The factor 1/2 occurs in the formula
because for anP undistorted triangle
R of unit area, |J | = 2 throughout the
triangle. Since wi = 1, we get dA = 1 when = 1. The area coordinates
and the weights for various orders of integration are presented in the following
table.
2 1 1
, ,0 1 (4.11)
2 2 3
0, 21 , 21 1
3
1
2
, 0, 21 1
3
1 1 1
3 , ,
3 3 3
0.5625
3 1 1
, ,
5 5 5
0.5208333..
1 3 1
, ,
5 5 5
0.5208333..
1 1 3
, ,
5 5 5
0.5208333..
The location of the Gauss points and the associated weights in Table (4.11)
are found as follows. If f (, ) = a0 + a1 + a2 , then using Eqn. (4.10), we
get
1
a0 + (a1 + a2 ) = w1 (a0 + a1 1 + a2 1 ),
3
which yields w1 = 1 and 1 = 1 = 1/3. If f (, ) = a0 + a1 + a2 + a3 2 +
98
a4 + a5 2 , then using Eqn. (4.10), we get
1 1 1
a0 + (a1 +a2 )+ (a3 +a5 )+ a4 = w1 (a0 +a1 1 +a2 1 +a3 12 +a4 1 1 +a5 12 )
3 6 12
+ w2 (a0 + a1 2 + a2 2 + a3 22 + a4 2 2 + a5 22 ).
1 1 1
a0 + (a1 +a2 )+ (a3 +a5 )+ a4 = w1 (a0 +a1 1 +a2 1 +a3 12 +a4 1 1 +a5 12 )
3 6 12
+w2 (a0 +a1 2 +a2 2 +a3 22 +a4 2 2 +a5 22 )+w3 (a0 +a1 3 +a2 3 +a3 32 +a4 3 3 +a5 32 ),
then we have more unknowns than equations. The Gauss points can now be
chosen to be symmetrically located, but as seen in Table (4.11), the choice
is not unique.
We have seen that for a three-node triangle, the stiffness matrix can be
constructed explicitly, and there is no need of numerical integration. For a
quadratic triangle,
B t CB |J| = f ( 2 , , 2).
Hence, a three point integration is exact when the triangle is an equilateral
triangle. As in the case of the quadrilateral elements, the integration formula
is approximate when the triangular element is distorted.
c = C(B u 0c ) + 0c ,
then the results obtained can be quite inaccurate since the process of dif-
ferentiation results in a loss of accuracy. For isoparametric elements, the
stresses are most accurate at Gauss points of a quadrature rule one order
lower than that required for full integration of the element stiffness. Such
points are known as Barlow points.
Consider the examples shown in Fig. 4.12. In example (a), the displace-
ments are of the form u = a, and v = 0, where a is a positive constant.
Thus, the shear strain xy is proportional to () yielding the variation shown
in the figure. Since the exact shear strain is zero, we see that the finite ele-
ment solution matches the exact solution at = 0 and = 0, which is also
the Gauss point location for a order-1 integration rule. Similarly, as shown
in Fig. 4.12b, for an 8-node (or even 9-node) element, the Barlow points (or
superconvergent points) are the Gauss-point locations for the 2 2 inte-
gration rule ( = = 1/ 3). These conclusions are true for rectangular
elements. For distorted elements, Gauss points may not be optimal locations,
but, nonetheless, they remain good choices.
To obtain more accurate stresses than those obtained from direct compu-
tation, the strategy used is to extrapolate the stress values from the Gauss
99
xy
P Exact
2L
0 L
P
(a)
xy Exact
V x
0 L 2L
(b)
r=1 =1
s=1
point to the nodes, and then average the values obtained from all the el-
ements sharing that node to obtain the nodal stresses. An example of an
extrapolation strategy is shown in Fig. 4.13. Since the element shown is a
four-node element, the 2 2 Gauss points are located at = = 1/ 3.
Let r and s be a scaled version of the - system such
that they have the
sameorigin but with r = s = 1 when = = 1/ 3. In other words, let
r = 3 and s = 3. The extrapolated stress components are given by (
is a typical stress component)
4
X
= Ni i ,
i=1
where i are the values of the stress components at the Gauss points, and
1
Ni = (1 r)(1 s).
4
100
Thus, the stress components
at point A in Fig. 4.13 are obtained by substi-
tuting r = s = 3 in the above formula. The four values obtained at each
node are averaged, and this mean value is taken as the nodal value. The
improved stress field is given by
n
X
= Ni (mean )i ,
i=1
where n is the number of nodes, and Ni are the displacement shape functions.
T = N1 T1 + N2 T2 + N3 T3 ,
where
1 y23 y31 y12
B= .
det J x32 x13 x21
101
Assuming that K is constant and isotropic over each element, i.e., K = kI,
the volume term of the stiffness matrix is
Z
(e)
K 1 = ke B t Bt dA = tke Ae B t B.
The surface area element is given by t ds = tl23 d. Hence, the surface term
of the stiffness matrix is
Z 1
(e)
K2 = hN t N tl23 d
0
0 0 0
htl23
= 0 2 1 .
6
0 1 2
We get
1
tQe Ae
f (1) = 1 if Q is constant,
3
1
2Q1 + Q2 + Q3
tAe
= Q1 + 2Q2 + Q3 if Q = N1 Q1 + N2 Q2 + N3 Q3 ,
12
Q1 + Q2 + 2Q3
0
= tQ0 N0 (0 , 0 ) = tQ0 0 if Q is a point source.
1 0 0
Assuming that the edge 2-3 constitutes part of q , and assuming that (q +
hT ) = constant, the load term due to the prescribed heat flux and convec-
tion is
Z 1
(2)
f = N t (q + hT )tl23 d
0
102
T = 180o C 5 T = 180o C 4
3
3
h = 50 W/m2o C q = 0 2 0.3 m
0.6 m q = 0 T = 25o C 1 q = 0
1 2
0.4 m
o
q = 0
T = 180 C
0.4 m
0
1
= (q + hT )tl23 1 .
2
1
As an example, consider the problem shown in Fig. 4.14 where the do-
mains conductivity is K = kI, where k = 1.5 W/m o C. Due to the symme-
try of the domain and boundary conditions, we can model only the upper half
of the domain as shown. Due to symmetry, the heat flux at the bottom edge
of the finite element analysis domain is q = T
n
= 0. On the right edge, since
T
there is no prescribed flux, we have n = h(T T ). The connectivity is
given by
11 2 3
25 1 3
35 3 4
while that due to the convection part (assuming that the convection is taking
103
place at edge 2-3 of the element) is
0 0 0
(e) hl23
Kh = 0 2 1 ,
6
0 1 2
104
12
z 11
10
u6
9 u5
8
7 u2 u4
5 6 u1 u3
4
r 1 2 3
c = C(c 0c ),
u = N u
N1 0 N2 0 N3 0 h it
= u1 u2 u3 u4 u5 u6 .
0 N1 0 N2 0 N3
r = r1 + r2 + (1 )r3 ,
z = z1 + z2 + (1 )z3 .
105
In the usual manner, we have
u u w w
r =J 1
; r =J 1
,
u u w w
z z
Note that det J = r13 z23 r23 z13 = 2A. The engineering strain vector is
given by
c = B u,
where
z23 z31 z12
0 0 0
|J | |J| |J |
r32 r13 r21
0 0 0
|J | |J| |J |
B=
r32
|J | z23 r13 z31 r21 z12
|J | |J| |J| |J | |J |
N1 N2 N3
r
0 r
0 r
0
The element stiffness matrix is given by
Z
K = 2 B t CBr drdz
(e)
Z 1 Z 1
= 2 B t CBr(2Ae ) dd
0 0
Z 1 Z 1
= 4Ae B t CBr dd,
0 0
106
Since the 2 factor is common in both, the stiffness and the force matrices,
one can cancel it out, or alternatively, carry out the formulation using a one
radian sector in the direction. An example of a axisymmetric problem is a
rotating flywheel with br = r 2 and bz = g.
For the four-node bilinear element, the strain-displacement matrix is
found in the same manner as in the plane stress/plane strain case. We have
u
r
rr
u
z
zz
= H w
, (4.12)
rz
r
w
z
u
where
1 0 0 0 0
0 0 0 1 0
H = .
0 1 1 0 0
0 0 0 0 1r
Also
u u
r
u u
z
w w
r = A , (4.13)
w w
z
u u
where
J 1 0 0
A= 0 1
0 .
J
0 0 1
Finally,
u
u
w
= Qu, (4.14)
w
u
107
where
N 0 N2, 0 N3, 0 N4, 0
1,
N1, 0 N2, 0 N3, 0 N4, 0
Q = 0 N1, 0 N2, 0 N3, 0 N4, .
0 N1, 0 N2, 0 N3, 0 N4,
N1 0 N2 0 N3 0 N4 0
B = HAQ.
Some terms in the K matrix are of the form 1/r. With Gauss quadrature,
these terms remain finite since there are no Gauss points at r = 0. For stress
computation, the indeterminate form = u/r = 0/0 arises for points on
the z-axis. Hence, it is convenient to calculate the stresses at the Gauss
points and extrapolate them to the axis. Another way is to use the fact that
= rr at r = 0. Thus, for r = 0, replace the row of the B matrix by
the rr row. The radial displacement u = 0 has to be prescribed at all nodes
that lie on the z-axis. Rigid body motion is restrained by prescribing w on
a single nodal circle.
108
Chapter 5
109
= a(uh , uh ) + a(eh , eh ) + 2a(uh , eh )
= a(uh , uh ) + a(eh , eh ), (5.1)
a(eh , eh ) a(eh + wh , eh + wh ).
110
y( x) H(x-y) x-y
1 (1-y)
y x y x y 1 x
d2 u
+ f = 0 on [0, 1], (5.2)
dx2
with the boundary conditions u(0) = u(1) = 0. The variational formulation
is Z 1 Z 1
du dv
dx = f v dx v H01 (I). (5.3)
0 dx dx 0
The solution to Eqn. 5.2 when f = y (x) (unit point load at x = y) is known
as the Greens function (denoted by G):
d2 G
+ y (x) = 0 on [0, 1], (5.4)
dx2
with G(0) = G(1) = 0. Integrating Eqn. 5.4, we get
G(x) + hx yi = c1 x + c2 ,
111
= 0,
where the last step follows from the error equation. Note that we have used
the fact that G belongs to Vh in the penultimate step. Thus,
uh (xi ) = u(xi ) i = 1, 2, . . . , n,
and we get the result that the finite element solution is exact at the nodes.
In some other kinds of one-dimensional problems, and in two and three-
dimensional problems, we do not, in general, have this property since the
Greens function does not necessarily belong to Vh .
a(u uh , u uh ) 0 as h 0,
a(uh , uh ) a(u, u) as h 0.
Physically, the above statement means that the strain energy calculated by
the finite element solution converges to the exact strain energy.
112
11
00
00
11
00
11
00
11
00
11
111
000
000
111
000
111
000
111
000
111
Figure 5.2: Element undergoing rigid-body displacement.
1. Within each element, the assumed field must contain a complete poly-
nomial of degree m,
a0 + a1 x + a2 y + a3 x2 + a4 xy + a5 y 2.
113
(bilinear) 1
(9 node) Q2 Q1
x y
CST
2 2
x xy y
quadratic triangle
3 2 2 3
x xy xy y
cubic triangle
4 3 2 2 3 4
x xy xy xy y
u = a1 + a2 x + a3 y + a4 x2 ,
v = a5 + a6 x + a7 y + a8 y 2 ,
u = a1 + a2 x + a3 y + a4 xy,
v = a5 + a6 x + a7 y + a8 xy,
114
The dependence of c on the material properties is detrimental when (almost)
incompressible materials are considered because c becomes very large, and
the order of convergence k results in good accuracy only at very small (im-
practical) values of h. The constant c depends on the type of element being
used. For example, the CST and Q4 elements which both belong to the
k = 1 group give different magnitudes of error for a given h, while the order
with which the error decreases as the mesh is refined is the same. Eqn. 5.5
gives in essence an error estimate for the displacement gradient (and hence
for the stresses and strains) because the primary contribution to the H 1 ()
norm is due to errors in the derivatives of displacements. The error in the
displacements is
Thus, the order of convergence for the displacements is one order higher than
for the strains.
These results are intuitively reasonable. If we think in terms of a Taylor
series analysis, a finite element of dimension h with a complete displacement
field expansion of order k can represent displacement variations upto that
order exactly. Hence, the local error inrepresenting arbitrary displacements
with a uniform mesh should be o hk+1 . The stresses are obtained
by differ-
entiating the displacements, and hence have error of o hk .
In the above convergence study, it is assumed that uniform discretiza-
tions are used (e.g., square elements in 2-D), and that the exact solution is
smooth. If the solution is not smooth, and a uniform mesh is used, the order
of convergence decreases. In practice, graded meshes are used with small
elements in the areas of stress concentration and larger elements away from
those regions. In general, a refined mesh is required in places where acute
changes in geometry, boundary conditions, loading, material properties or
solution occur. A good mesh is one for which the error is almost equal for
all the elements, and below a certain tolerance.
The method just discussed where we increase the number of elements
keeping the interpolation functions fixed is known as h-adaptivity. An al-
ternative method of increasing the accuracy of the solution is the p method
where the order of the interpolation functions in high-error elements is in-
creased using hierarchical shape functions.
A very high rate of convergence in the solution can be obtained if we
increase the number of elements and at the same time increase the order of
interpolation functions. This approach is known as the h-p method. In all
of the above approaches, the mesh grading should be such that the error in
each element is almost the same.
Mesh distortion and numerical integration will not reduce the order of
convergence provided the distortion is reasonable and full numerical integra-
tion is used.
115
(a) (b)
L
1 2 B
L
incompatibility C
(c)
(d)
116
Figure 5.5: Mesh grading using isoparametric elements.
117
hm
hm
m
m
Aspect ratio distortion Parallelogram distortion
hm
m
m
hm
Angular distortion
Curved-edge distortion
Midnode distortion
118
10
P
P 2
B
C A
A
xB=0.096 vA =0.091 xC=0.727 vA =0.682
C A
xC=0.301 vA =0.494
119
P P
B B
B
B vA B vA
B vA
2x2 1.0 0.968 0.051 0.362 -0.048 0.430
8-node 0.048 0.161 0.050 0.221
3x3 1.129 0.93
1.006 1.125 1.109 0.958 0.955
9-node 2x2 1.00
0.737
3x3 1.141 0.954 0.687 0.791 0.705
b M M
u u
u = u; v = 0,
where Ni are the usual bilinear functions. The above element represents
bending exactly, but now the compatibility of displacements at the element
edges is lost. Such elements are called incompatible or nonconforming ele-
ments.
The results obtained using incompatible elements are of high quality.
However, the exact potential energy of the system is not necessarily an upper
bound on the calculated potential energy. Hence, monotonic convergence is
not ensured. But we can establish conditions so that there is at least non-
monotonic convergence. The conditions are
The element completeness condition must always be satisfied, i.e., it
should be able to represent rigid body modes and the constant strain
states
As the finite element mesh is refined, each element should approach a
constant strain condition. Hence, a assemblage of incompatible finite
elements should be able to represent constant strain conditions. Note
that this condition is on an assemblage of elements and not on a single
element.
To test the second condition, the patch test has been proposed. In this test
a patch of elements is subjected to nodal loads consistent with constant xx ,
yy and xy in a two-dimensional problem (or all the six stress components in
a three-dimensional one). If for any patch of elements, the element stresses
actually represent the constant stress conditions, then the element passes the
patch test. If an incompatible element passes the patch-test then convergence
is ensured (although it may not be monotonic). In the case of a compatible
element, we know that the convergence is monotonic. In such a case, the
patch test can be used as a debugging aid.
Two examples of the patch test are shown in Fig. 5.10. The roller supports
are provided since a fixed support would prevent the Poisson contraction in
the y-direction. Actually, we need to test all possible patches. But testing
just one patch is what is practiced. This practice is generally found to be
quite reliable.
One can design variations of the patch test. For example, if an element
fails to display the constant stress state in a patch of elements, but displays
the constant stress state as the mesh is repeatedly subdivided, then a weak
patch test can be designed with a larger number of elements in the assem-
blage being tested. Similarly a higher-order patch test can be designed for
121
P
P/2
2P
P/2
P
122
Chapter 6
Eigenvalue and
Time-Dependent Problems
123
or
d dU
E U = 0, (6.3)
dx dx
where 2 . Eqn. 6.3 is an eigenvalue problem which involves finding
and U.
For constant E and , the solution is given by
r r
U(x) = c1 sin x + c2 cos x,
E E
yielding " r r #
u(x, t) = c1 sin x + c2 cos x eit .
E E
The constants c1 , c2 and are determined from the initial and boundary
conditions.
Now consider the parabolic partial differential equation
T T
c k = Q. (6.4)
t x x
We seek a homogeneous solution to the above equation in the form
T = U(x)et . (6.5)
v = U(x)eit ,
124
The study of buckling also leads to an eigenvalue problem. The equation
of equilibrium of a beam subjected to an axial force P is
d2 d2 v d2 v
EI + P = 0,
dx2 dx2 dx2
2u
= + b.
t2
The virtual work principal is derived in the usual way. Starting from
2u
Z Z
+ b 2 v d + (t t) v d = 0 v.
t t
2u
Z Z Z Z
2 v d + bv d + tv d : (v) d = 0 v. (6.6)
t t
We shall use Eqn. 6.6 to develop the finite element formulation. Just for the
sake of completeness, however, we shall derive Hamiltons principle which we
had used in Section 1.2.5.
Assuming the existence of a strain-energy density function W for the
body so that
W
= ,
the above expression can be written as
2u
Z
2 v d + (1) = 0, (6.7)
t
where Z Z Z
= W d b u d t u d.
t
Now we enforce Eqn. 6.7 which holds at every instant of time t in a weak
sense with respect to time:
Z t2 Z 2 Z t2
u
2 v d dt + (1) dt = 0,
t1 t t1
125
which on interchanging the order of integration yields
Z Z t2 2 Z t2
u (1)
2 v dt d + dt = 0.
t1 t t1
u
Assuming that u is prescribed or t
= 0 at t1 and t2 , Eqn. 6.8 reduces to
t2 t2
(1) u u
Z Z Z
d dt + (1) dt = 0.
t1 2 t t t1
where
u u
Z
T = d
2 t t
is the kinetic energy of the body at time t. Alternatively, we can write the
above equation as Z t2
(1)
L dt = 0,
t1
where the shape function N are functions of the space variables, and the
vectors u are functions of time. It may appear from our semi-discrete ap-
proximate form given by Eqn. 6.9 that this approximation is valid only when
the exact solution is separable in x and t. This is not so. Even nonseparable
solutions can be approximated by our separable approximation. Substituting
Eqn. 6.9 in Eqn. 6.6, we get
+ K u = f ,
M u (6.10)
126
where
Z
M= N t N d,
Z
K= B t CB d,
Z h i Z
f= t t 0 t 0 t t
N b + B Cc B c B C B u N N u d + N t t d.
t
Since any system is not perfectly elastic, the effects of energy dissipation
have to be incorporated in Eqn. (6.10). These effects are incorporated by
modifying Eqn. (6.10) to
+ C u
M u + K u = f , (6.11)
where C is the damping matrix. Eqn. 6.11 is solved subject to the initial
conditions u(0) = u0 and u(0) 0.
= u
The damping matrix, C, is positive definite. To see this in the purely
mechanical theory where thermal effects are ignored, consider the first law
of thermodynamics
Z
d 1
Z Z
v v + e d = t v d + b v d.
dt 2 t
since the kinetic energy has the same value at times t1 and t1 + T , and since
the internal energy increases due to dissipation (the contribution to e due to
the elastic strain energy, however, is the same at times t1 and t1 + T ). Now
taking the dot product of the finite element equations
= M u
f u + C u
+ K u u
1 d
+ u K u + u
C u.
= u M u
2 dt
Integrating the above equation between the limits t1 and t1 + T and using
the fact that the kinetic and strain energies are the same at times t1 and
t1 + T , we get Z t1 +T Z t1 +T
C u
u dt = dt > 0.
f u
t1 t1
t1 and T , we conclude
Since the above equation holds for all choices of u,
that C is positive definite.
127
There is no systematic technique to generate the damping matrix unlike
the mass and stiffness matrices. A simple but popular model that is usually
used is Rayleigh or proportional damping in which we assume
C = M + K,
Assume a solution for u of the form u = ueit , where u is the set of constant
values at the nodes called as the modal vector, and is known as the natural
frequency. Substituting this solution in Eqn. 6.12, we get
(K 2 M )u = 0.
For a nontrivial solution to the above equation, it is necessary that the de-
terminant of the coefficients of u be zero, i.e.,
det(K 2 M ) = 0.
The eigenvalues are real. To see this, let and the corresponding mode
shape be complex-valued. Then
u K u = 2 u M u,
u K u = ( 2 ) u M u ,
where denotes complex conjugate. Subtracting the first equation from the
second, and using the symmetry of K and M , we get
u M u[ 2 ( 2 ) ] = 0.
where u = N u. Thus,
2 = ( 2) ,
i.e., 2 is real-valued. Since now uK u = 2 uM u, and since M is positive
definite and K is positive semi-definite, we have 2 0. If the order of K
and M is n, then there are n natural frequencies and correspondingly n
natural mode shapes or eigenvectors.
Assuming that all eigenvalues are distinct, the eigenvectors are orthogonal
in the following sense:
ui K uj = 0, i 6= j,
ui M uj = 0, i 6= j.
128
The proof is as follows. For the i th and j th modes, we have
K ui = i2 M ui ,
K uj = j2 M uj .
Hence,
0 = (i2 j2 )utj M ui .
For i 6= j, i 6= j , we get
utj M ui = 0.
Hence, from Eqn. 6.13 utj K ui = 0. For i = j, we get
uti M ui = EM ,
uti K ui = i2 EM ,
uti M ui = 1,
uti K ui = i2 .
v t Kv
Q(v) = ,
vtM v
where v is an arbitrary vector. One of the important properties of the
Rayleigh quotient is
12 Q(v) n2 ,
where the frequencies 1 , 2 ,..., n are assumed to be ordered such that
1 2 . . . n .
129
Substituting this expansion in the expression for Q(v), and using the orthog-
onality property of u, we get
12 12 + 22 22 + + n2 n2
Q(v) =
2 + 22 + + n2
1 2 2
2 2 2 2 n
1 + 2 1 + + n 1
= 12
12 + 22 + + n2
12 .
Similarly,
2 2
2 1 2
1 n
+ 22 n
++ n2
Q(v) = n2
12 + 22 + + n2
n2 .
12 = minn Q(v),
v
n2 = maxn Q(v).
v
hn n2 max(max
2
)e ,
e
2
where (max )e is the maximum eigenvalue of element e found by solving the
eigenvalue problem for that element:
for the various element types that we have studied so far. We shall assume
that = constant.
130
Hence, the consistent element mass matrix is
Z
(e)
M = e N t N A dx
e
e Ae le 1 t
Z
= N N d
2 1
e Ae le 2 1
= .
6 1 2
Truss element
If u1 -u4 represent the degrees of freedom of a truss element, then
u
1
u N 0 N2 0 u2
= 1 ,
v 0 N1 0 N2 u3
u4
thus yielding
N1 0 N2 0
N = ,
0 N1 0 N2
where N1 = (1 )/2 and N2 = (1 + )2. Substituting in Eqn. 6.15, we get
2 0 1 0
0 2 0 1
e Ae le
M (e) = .
6 1
0 2 0
0 1 0 2
Beam element
Using the Hermite shape function matrix, i.e.,
le le
N = N1 N2 N3 N4 ,
2 2
131
where
1
N1 = (2 3 + 3),
4
1
N2 = (1 2 + 3 ),
4
1
N3 = (2 + 3 3),
4
1
N4 = (1 + 2 + 3 ).
4
Substituting in Eqn. 6.15, we get
Z 1
(e) e Ae le
M = N tN d
1 2
156 22le 54 13le
2 2
22le 4le 13le 3le
e Ae le
=
420 54 13le 156 22le
2 2
13le 3le 22le 4le
Frame element
In the body coordinate system, the mass matrix is a combination of the bar
element and the beam element:
2a 0 0 a 0 0
0 156b 22le b 0 54b 13le b
2 2
0 22l b 4l b 0 13l b 3l b
(e)
e e e e
Mb =
a 0 0 2a 0 0
0 54b 13le b 0 156b 22le b
2 2
0 13le b 3le b 0 22le b 4le b
where a = e Ae le /6 and b = e Ae le /420. In the global coordinate system,
the element mass matrix is given by
(e)
M (e) = Lt M b L,
where
l m 0 0 0 0
m l 0 0 0 0
0 0 1 0 0 0
L=
.
0 0 0 l m 0
0 0 0 m l 0
0 0 0 0 0 1
132
CST element
The shape function matrix is given by
N1 0 N2 0 N3 0
N = ,
0 N1 0 N2 0 N3
ZA
= 2e (N1 r1 + N2 r2 + N3 r3 )N t N dA
A
4r1
+ 2r 0 2r r33 0 2r r2
0
3 3
4r1
0 + 2r 0 2r r33 0 2r r2
3 3
2r r33 0 4r2
+ 2r 0 2r r1
0
e Ae 3 3
= ,
10 0
2r r33 0 4r2
+ 2r 0 2r r1
3 3
2r r32 0 2r r31 0 4r3
+ 2r 0
3
0 2r r32 0 2r r31 0 4r3
3
+ 2r
Q4 element
The shape function matrix is given by
N1 0 N2 0 N3 0 N4 0
N = ,
0 N1 0 N2 0 N3 0 N4
133
2
A=(1+ /40) cm 2
A=1 cm 2
100 cm 80 cm
Example
Consider the structure shown in Fig. 6.1 with the material properties E =
60 109 Pa and = 7500 kg/m3 . Using two linear elements to model the
structure, the mass and stiffness matrices are
0.0025 0.00125 0
2
M = 10 0.00125 0.0073 0.0042
0 0.0042 0.0128
6 6 0
6
K = 10 6 38.5 32.5
0 32.5 32.5
8 6
On solving the
above equation, we get 1 = 1.3035
10 and 2 = 1.97 10 .
Thus, 1 = 1 = 11417.3 rad/s and 2 = 2 = 1404.854 rad/s. The
134
0.0006169
0.000555
0.000763
100
180
0.001175
T = T on T ,
h(T T ) + (kT ) n = q on q ,
T (x) = T0 (x) at t = 0.
135
T
Z Z Z Z
t
c v d + v kT d + vhT d = vQ d+
t q
Z
v(q + hT ) d v VT ,
q
subject to the initial condition T (x, 0) = T0 (x). The finite element formu-
lation is carried out similar to that in the elasticity case. Discretizing the
temperature as X
T = N (x)u(t) + N (x)u(t), (6.16)
and substituting in the variational formulation, we get
M u + Ku = f
subject to u(0) = u0 ,
Note the similarity of the form of M with the form of the mass matrix
(also denoted by M ) in the dynamics formulation. In particular it is sym-
metric and positive definite. In practice, diagonal or lumped mass matrices
are used instead of the consistent mass matrix derived above, because they
lead to economical time-integration schemes (so called explicit methods).
The eigenvalue problem can be derived by assuming
u = uet ,
(K M )u = 0,
which is similar in form to the eigenvalue problem for the dynamics problem.
Example:
Consider a plane wall initially at uniform temperature T0 , which is suddenly
exposed to a fluid at temperature T . The governing equation is
2T T
2
= ,
x t
where = k/(c) is the diffusion coefficient. The initial condition is T (x, 0) =
T0 . Consider two sets of boundary conditions:
136
2. T (0, t) = T , kA T
x
+ hA(T T )x=L = 0.
To find the solution, we define some nondimensional parameters:
x t T T
x = ; t = 2 ; u = .
L L T0 T
The differential equations and boundary conditions in terms of these param-
eters are (dropping the bars on x and t for notational convenience)
2 u u
+ = 0,
x2 t
1. u(0, t) = 0, u(1, t) = 0, u(x, 0) = 1.
2. u(0, t) = 0, u
x
+ Hux=1 = 0, u(x, 0) = 1, where H = hL/k.
Assume u = et U(x). Substituting in the governing differential equation,
we get
d2 u
2 U = 0.
dx
Solving the above equation, we get
h i t
u = c1 sin x + c2 cos x e .
1. For the first set of boundary conditions, we get c2 = 0, = n. Thus,
the exact solution is
X
2 2
u= [cn sin nx] en t .
n=1
The constants cn can be determined from the initial condition, and the
orthogonality of the sine functions:
X
1= cn sin nx
n=1
137
for 2 linear elements, 1 = 12.
for 1 quadratic element, 1 = 10.
exact, 1 = 2 .
2. Assuming
H = 1, the second set of boundary conditions yields Un (x) =
sin n x, where n satisfies
p p
1 + n cot n = 0.
M u + Ku = f , (6.17)
subject to the initial condition u(0) = u0 . First we state the exact solution
to these set of equations. Multiplying Eqn. (6.17) by eM Kt M 1 , we get
1
d h M 1 Kt i
u = eM Kt M 1 f .
1
e
dt
138
Integrating this equation either between the limits [0, t] or [t, t + t], we get
Z t
M 1 Kt
eM K(t) M 1 f () d,
1
u(t) = e u0 + (6.18a)
0
Z t+t
M 1 Kt
eM K(tt) M 1 f () d. (6.18b)
1
u(t + t) = e u(t) +
t
Method
The problem is to determine un+1 and v n+1 given un and v n . The initial
value v 0 is determined from
M v 0 = f 0 Ku0 .
139
We now show how a form close to Eqn. (6.22) can be obtained as an
approximation to the exact solution given by Eqn. (6.18b). In the derivation,
we will use the property that eA+B = eA eB if and only if A and B commute.
First multiply Eqn. (6.18b) by eM Kt to get
1
Z t+t
M 1 Kt (1)M 1 Kt
eM K(t(1)t)
M 1 f () d.
1
e un+1 = e un +
t
eM Kt
I + tM 1 K,
1
e(1)M Kt
I (1 )tM 1 K,
1
eM Kh
1
I,
where in the last approximation, we have retained only the first term to get
a linear approximation in t in the final result. Substituting Eqn. (6.24) and
the above approximations into Eqn. (6.23), and carrying out the integration,
we get
1
I + tM 1 K un+1 = I (1 )tM 1 K un + tM 1 f n + f n+1 .
2
Multiplying the above equation by M , we obtain a form close to Eqn. (6.22).
un u(tn ) as t 0.
140
Modal reduction to SDOF form
The essential property used in reducing to SDOF form is the orthogonality
of the eigenvectors of the associated eigenvalue problem
(K i M )i = 0; i = 1, 2, . . . , n
with no sum on i. The eigenvectors i |ni=1 constitute a basis for n , i.e., any
element of n can be written as a linear combination of i |ni=1 . Therefore,
n
X
u(t) = uj (t)j ,
j=1
n (6.26)
X
u(t) = uj (t)j ,
j=1
where
ui (t) = ti M u(t),
ui (t) = ti M u(t).
ui + i ui = ti f (t), i = 1, 2, . . . , n.
ui (0) = ti M u(0) = ti M u0 .
141
M u + Ku = f
temporal (M + tK)un+1 = (M (1 )tK)un + tf n+
u(0) = u0
discretization u0 given
modal modal
decomposition decomposition
u + u = f
temporal (1 + t)un+1 = (1 (1 )t)un + tfn+
u(0) = u0
discretization u0 given
n
X
un+1 = un+1(j) j ,
j=1
where
un(i) = ti M un ,
un+1(i) = ti M un+1 .
142
The proof is as follows. The errors in the original and SDOF system at time
tn are given by
e(tn ) = u(tn ) un ,
ei (tn ) = ui (tn ) un(i) .
Hence,
n
X
t
e (tn )M e(tn ) = (ei (tn )i )t M (ej (tn )j )
i,j=1
X n
= ei (tn )ej (tn )ti M j
i,j=1
X n
= ei (tn )ej (tn )ij
i,j=1
X n
= [ei (tn )]2 .
i=1
Hence,
et (tn )M e(tn ) 0 |ei (tn )| 0 i.
Since M is assumed to be positive definite,
Stability
If u now denotes a perturbation of the solution, then noting that Eqn. (6.27)
is linear, we get
u + u = 0.
Since the solution of this equation is given by cet , we have u(tn ) = cetn
and u(tn+1 ) = cetn+1 . Thus,
For the solution to remain stable, i.e., the condition that the perturbation
not blow up, we need
|u(tn+1 )| < |u(tn )| , > 0,
(6.29)
u(tn+1 ) = u(tn ), = 0.
These are the conditions that we wish to mimic in the temporally discrete
case.
If un and un+1 denotes the perturbations in the finite element solution at
times tn and tn+1 , then using Eqn. (6.28), we have
(1 + t)un+1 = (1 (1 )t)un .
143
Note that (1 + t) > 0 since and t are both greater than zero. We
can write the above equation as
un+1 = Aun , (6.30)
where
1 (1 )t
A= ,
1 + t
is the amplification factor. To mimic Eqn. 6.29, we need
|un+1| < |un | , > 0,
un+1 = un , = 0.
From the definition of A, we see that the second condition is automatically
satisfied. The first condition implies |A| < 1, or 1 < A < 1, i.e.,
1 (1 )t
1 < < 1.
1 + t
The right-hand inequality is automatically satisfied. The left-hand inequality
is satisfied whenever 0.5. If < 0.5, then
2
t < .
1 2
An algorithm for which stability imposes a time-step restriction is called
conditionally stable. An algorithm for which there is no time-step restriction
imposed by stability is called unconditionally stable. Even if the amplifica-
tion factor, A, is slightly greater than 1, disastrous growth can occur. For
example, for |A| = 1.01 and n = 1000, we have
un
= An = 1.011000 = 2.09 104 .
u0
Note the following points:
1. In the conditionally stable case, the stability condition
2
t < ,
(1 2)
must hold for all modes i , i = 1, 2, . . . , n. Hence, the greatest i , i.e.,
n imposes the most stringent restriction on the time step, i.e.,
2
t < .
(1 2)n
In heat transfer problems, it can be shown that n = o (1/h2 ), where
h is the typical element size. Thus, we need t < ch2 , where c is
some constant. This, is a very severe constraint on the step size; thus,
unconditionally stable algorithms are generally preferred.
2. For = 0.5 and t 1, A 1. Thus, high modal components
will behave like (1)n (sawtooth pattern). These spurious modes are
filtered out by reporting step-to-step averages , (un+1 + un )/2.
144
Convergence
The temporally discrete model problem can be written as
where
tfn+
Ln = .
1 + t
If we replace un and un+1 by the exact values, we get
= (1 2)o (t) + o t2 .
Thus, the trapezoidal rule ( = 0.5) is the only member of the family of
methods that is second-order accurate.
We have the following result:
Theorem: Consider Eqns. 6.31 and 6.32. Let tn be fixed (n, and hence
t = tn /n are allowed to vary). Assume that the following conditions hold
1. |A| 1 (stability)
Thus,
145
The first term on the right vanishes since e(0) = u0 u(0) = 0. Thus,
n1
X
|e(tn )| = t Ai (tn1i )
i=0
n1
X
t |A|i | (tn1i )|
i=0
n1
X
t | (tn1i )| (by stability)
i=0
tn max | (t)| (t = tn /n)
tn ctk (consistency)
Hence, e(tn ) 0 as t 0, and furthermore the rate of convergence is k,
i.e., e(tn ) = o tk . The maximal rate of convergence is 2.
The theorem just proved is a particular example of the Lax equivalence
theorem which states that consistency plus stability is necessary and sufficient
for convergence.
146
3. The approximation to u(t) is synthesized from the modes, viz.,
nX
modes
u(t) ui (t)i .
i=1
The calculation in the above equation must be performed for each time
t at which the solution is required.
1. they are efficient if many analyses of the same configuration are re-
quired.
147
Z t+t
M
1
Kt
1 1
u(t + t) = e u(t) + eM K(tt)
M f () d, (6.35b)
t
where
1 0 I u0
M K = , u0 = .
M 1 K M 1 C v0
We now discuss the modal analysis and direct methods.
Ki = i2 M i
t M = I,
t K = ,
where is a matrix with the eigenvectors i2 along the diagonal and all
other elements zero. Since the vectors i form a basis, we can write the
displacement vector, u as
u = u.
Since is independent of time, we can write Eqn. (6.34) as
M u + Cu + Ku = f .
u + t Cu + u = f ,
148
which are to be solved subject to the initial conditions u(0) = t M u0 and
u(0) = t M u 0 . Based on the analogy with a sping-mass-damper system,
each diagonal element of is conventionally written as 2i i , where i > 0
is known as the modal damping parameter. The damping ratio = 1 marks
the transition between oscillatory and nonoscillatory response, and is hence
known as the critical damping ratio. Damping ratios are found experimen-
tally by observing the vibratory response of the structure. Typically, ranges
between 0.005 to 0.15.
If the modal damping parameters (and hence ) is known, then the damp-
ing matrix can be constructed explicitly as
C = M t M .
One can easily verify from the above equation that t C = . However,
since computing the matrix is quite an expensive proposition, one usually
assumes Rayleigh or proportional damping, i.e.,
C = M + K,
+ i2 = 2i i i = 1, 2, ..., n. (6.36)
Obviously the above equation cannot hold, in general, for all values of i
since there are only two undetermined constants and . Hence, the above
equations are solved for two given damping ratios (which are determined
experimentally) corresponding to two distinct natural frequencies 1 and 2 .
We get
21 2 (1 2 2 1 )
= ,
22 12
2(2 2 1 1 )
= .
22 12
Once and are determined, the remaining modal damping parameters are
determined from Eqn. (6.36), though these will in general, not match with
the actual damping parameters of the system. Note that in this approach C
is banded, whereas in other approaches, it might be fully populated.
One usually chooses 1 to be the lowest natural frequency and 2 to be
the maximum frequency of interest in the loading or response. Choosing 1
to be the lowest natural frequency guarantees that C is positive definite. To
see this, let a be some arbitrary vector. We want to prove that at Ca > 0.
Since the columns of act as a basis, we can express a as b, where b is
another vector. Thus, if b (b1 , b2 , ..., bn ) then
at Ca = bt t (M + K)b
= bt (I + )b
149
= ( + 12)b21 + + ( + n2 )b2n
> 0,
+ n2 + n1
2
+ 12 = 21 1 > 0.
(K i2 M )i = 0 i = 1, 2, . . . , neq ,
where 0 1 2 . . . n and
ti M j = ij
ti Kj = i2 ij (no sum on i).
ui + 2i i ui + i2 ui = fi 1 i nmodes ,
where ui (0) = u0(i) and ui (0) = u0(i) . Solutions to the above equations
can be obtained easily if f (t) is a simple function of time, or by a step-
by-step method using a very small time step. Since the problem in step
2 is of a scalar nature, the computational cost in step 2 is negligible
compared to step 1.
u(t) ui (t)i .
i=1
The calculation in the above equation must be performed for each time
t at which the solution is required.
d
Z Z Z
v dV = t dS + b dV,
dt V S V
150
d
Z Z Z
(X v) dV = X t dS + X b dV,
dt V S V
d vv
Z h i Z Z
+ W () dV = t v dS + b v dV,
dt V 2 S V
151
which proves that the linear momentum is conserved.
Now choose u = c X over the entire domain, where c is a constant
vector. If W is the skew tensor whose axial vector is c, then we have u =
W X, so that u = W , again resulting in (u ) = 0. Using the property
(p q) r = p (q r), Eqn. (6.39) reduces to
Z
c X (v n+1 v n ) dV = 0,
V
Note that the choices for u made above fall within the finite element space
for u, and hence the conservation properties hold in the fully-discrete setting.
By eliminating v n+1 from Eqns. (6.37) and (6.38), we get
2M K 2 2 1 1
+ un+1 = 2 M un + M v n K un + f n + f n+1 . (6.41)
t2 2 t t 2 2
Using the initial displacement and velocity vectors, u0 and v 0 , we first solve
for the nodal displacement vector ut1 at time t1 using Eqn. (6.41). Next,
we substitute for ut1 in Eqn. (6.38) to find v t1 . Using ut1 and v t1 , we find
(ut2 , v t2 ), and so on. Thus, starting from the initial displacement and velocity
fields, we march forward in time, until the time instant at which the response
is desired is reached.
Now consider a nonzero damping matrix C, which is assumed to be
positive semi-definite (e.g., C = M + K). Using the approximation
152
(un+1 un )/t in Eqn. (6.34), we now get instead of Eqn. (6.40) the
u
relation
1 1
Z Z
[v n+1 v n+1 v n v n ] dV + [n+1 : Cn+1 n : Cn ] dV
2 V 2 V
1
+ (un+1 un ) C(un+1 un ) = 0,
t
which by virtue of the positive definiteness of C shows that
The constants, and , determine the stability and accuracy of the algo-
rithm. The implementation to find un+1 , v n+1 and an+1 given un , v n and
an is described next.
6.4.3 Implementation
Define predictors
t2
un+1 = un + tv n + (1 2)an , (6.46)
2
v n+1 = v n + (1 )tan . (6.47)
Using the above relations, Eqns. 6.44 and 6.45 can be written as
153
v n+1 = v n+1 + tan+1 . (6.49)
M a0 = f Cv 0 Ku0 ,
After finding an+1 , we use Eqns. 6.48 and 6.49 to obtain un+1 and v n+1 ,
respectively.
C = aM + bK,
then the symmetry of M and K make the decomposition shown in Fig. 6.4
possible. The parameter in Fig. 6.4 is the undamped frequency of vibration,
while is known as the damping ratio, and is given by
1 a
= + b .
2
The SDOF model problem can be written as
y n+1 = Ay n + Ln ,
154
M u + C u + Ku = f M an+1 + Cv n+1 + Kun+1 = f n+1
temporal un+1 = un + tv n + t2
[(1 2)an + 2an+1 ]
u(0) = u0 2
discretization v n+1 = v n + t [(1 )an + an+1 ]
u(0) = v 0
u0 , v 0 given
modal modal
decomposition decomposition
where 1/2
( 0.5) + 2
+ 2( 0.5)2
crit =
,
2
is known as the critical sampling frequency. The stability condition must be
satisfied for each mode. Hence, the maximum natural frequency n is critical
for determining the time step. We have
crit
t
n
As usual n is bounded by the maximum frequency of the individual elements.
Observe that when = 0.5, viscous damping has no effect on the stability,
since we have 0.5
crit = .
2
When > 0.5, the effect of viscous damping is to increase the critical time
step of conditionally stable Newmark methods. Thus, the undamped critical
frequency (/2 )0.5 serves as a conservative estimate when an estimate
of the modal damping coefficient is not available.
155
6.4.5 Special cases of the Newmark method
Method Type Stability condition Order of
accuracy
1 1
Average acceleration implicit 4 2
unconditional 2
(trapezoidal rule)
1 1
Linear acceleration implicit 6 2
crit = 2 3 2
1 1
Fox-Goodwin implicit 12 2
crit = 6 2
(royal road)
1
Central difference explicit 0 2
crit = 2 2
156
where y(tn ) = [u(tn ) u(tn )]t . Using an analysis similar to that for parabolic
problems, we get
(t) = o tk
t [0, T ],
where k = 2 if = 0.5 and k = 1 otherwise.
The stability of the Newmark methods is determined by the properties
of the amplification matrix. The error equation is given by
n1
X
n
e(tn ) = A e(0) t Ai (tn1i ).
i=0
Once again, stability plus consistency implies convergence. The rate of con-
vergence is k.
Let i (A) denote the eigenvalues of A. The modulus of i (A) is written
as 1/2
|i (A)| = i (A)i (A) ,
where i (A) is the complex conjugate of i (A). The spectral radius of A,
(A) is defined by
(A) = max |i (A)| .
i
1. (A) 1
The above conditions define a spectrally stable A. Since the stability condi-
tion depends only on the eigenvalues of A, it can be expressed in terms of the
principal invariants of A. Thus, is given by the solution to 2 I1 +I2 = 0,
i.e., p
I1 I12 4I2
= ,
2
where
I1 = tr A = A11 + A22 ,
I2 = det A = A11 A22 A12 A21 .
157
6.4.8 Superconvergence in one-dimensional problems
Consider the one-dimensional wave equation
2u E 2u 2
2 u
= = c ,
t2 x2 x2
p
where c = E/ is the characteristic velocity or the speed of sound. t =
h/c is called the characteristic time step. It is equal to the transit time for a
wave moving at speed c to traverse one element.
The element stiffness and mass matrices are given by
EA 1 1 0.5 r r
K (e) = ; M (e) = Ah .
h 1 1 r 0.5 r
2. lumped mass (r = 0)
h 1 0
M (e) =
2 0 1
158
6.4.9 Time estimates for some simple finite elements
Consider the two-node linear rod element. If we assume a lumped mass
matrix then
EA 1 1 Ah 1 0
K (e) = ; M (e) =
h 1 1 2 0 1
4c2
2 = or = 0.
h2
Thus, max = 2c/h. For = 0, = 0.5,
2 h
tcritical = = = tcharacteristic .
max c
For a consistent mass matrix max = 2 3c/h, which results in a reduced
critical time step:
h
t .
3c
In general, consistent-mass matrices tend to yield smaller critical time steps
than lumped-mass matrices.
For the heat conduction problem with = 0, the condition t 2(1
2) implies that
2
t .
max
2
Since max = max = 4k/ch2 , we get
ch2
t .
2k
For a 3-node quadratic element,
1 0 0 7 8 1
(e) hA EA
M lumped = 0 4 0 ; K (e) = 8 16 8 ,
6 h
0 0 1 1 8 7
we get max = 2 6c/h, yielding
h
t .
6c
159
Thus, the allowable time step is about 0.4082 that for linear elements with
lumped mass.
For a four-node quadrilateral element
max cd g, (6.54)
p
where cd = ( + 2)/ is the dilatational wave velocity, and are Lame
constants, and g is a geometric parameter. For example, for a quadrilateral
2 4
4 XX
g= 2 Bij Bij ,
A i=1 j=1
For a rectangular element with sides h1 and h2 , the above formula reduces
to
1
t 1/2 .
1 1
cd h 2 + h 2
1 2
160
Appendix A
For a beam along the y-axis, interchange I and J in the above matrix.
161
The stiffness matrices for beams along the x and y-axes are given by
1 1 1 1
0 0
L 2 L 2
0 0 0 L 0
L
1
2 0 L4 + L 1
2
0 L
4
L
K x = GAk 1
,
1 1 1
0 0
L 2 L 2
0 0 0 0
L L
12 0 L
4
L
1
2
0 L
4
+
L
1 1
L 2
0 L1 1
2
0
1 L
+
0 21 L
0
2 4 L 4 L
0 0 L
0 0 L
K y = GAk .
1 12 0 1
12 0
L L
1 L
0 21 L
+
0
2 4 L 4 L
0 0 L 0 0
L
Quadratic element:
The mass matrix for a beam along the x-axis is
4A 0 0 2A 0 0 A 0 0
0 4J 0 0 2J 0 0 J 0
0 0 4I 0 0 2I 0 0 I
2A 0 0 16A 0 0 2A 0 0
L
M= 0 2J 0 0 16J 0 0 2J 0 .
30
0 0 2I 0 0 16I 0 0 2I
A 0 0 2A 0 0 4A 0 0
0 J 0 0 2J 0 0 4J 0
0 0 I 0 0 2I 0 0 4I
The mass matrix for a beam along the y-axis is obtained by interchanging I
and J in the above equation.
162
The stiffness matrices for beams along the x and y-axes are given by
7 1 8 2 1 1
0 0 0
3L 2 3L 3 3L 6
7 8
0 0 0 3L 0 0 0
3L 3L
1 7 L 2 8 L 1 L
2 0 3L
+9 3
0 3L + 9 6 0 3L
18
8 2 16 8 2
3L 0 0 0 0
3 3L 3L 3
K x = GAk 0 3L 8 16 8
0 0 0 0 3L 0 ,
3L
2 8 L 16 4L 2 8 L
3 0 3L + 9 0 0 + 9 0 3L + 9
3L 3
1 1 8 2 7 1
0 6 3L 0 0
3L 3 3L 2
8 7
0 3L
0 0 3L 0 0 3L
0
1 L 2 8 L 1 7 L
6
0 3L
18 3 0 3L + 9 2
0 3L
+9
7 1 8 2 1 1
0 0 0
3L 2 3L 3 3L 6
1 7 L 2 8 L 1 L
+9 0 3 3L + 9 0 18 0
2 3L 6 3L
7 8
0 0 0 0 0 0
3L 3L 3L
8
3L 23 0 16
0 0 3L 8 2
0
3L 3
2 8 L 16 4L 2 8 L
K y = GAk 3 3L + 9 0 0 + 9 0 3 3L + 9 0 .
3L
8 16 8
0 0 3L 0 0 0 0
3L 3L
1 1 8
0 3L 23 0 7
12 0
3L 6 3L
1 L 2 8 L 1 7 L
6 18 0 3L + 9 0 2 +9 0
3L 3 3L
8 7
0 0 3L
0 0 3L 0 0 3L
163
Bibliography
[2] Shames, I. H. and C. Dym, Energy and Finite Element Methods in Struc-
tural Mechanics, Hemisphere Publishing Co., 1985.
[3] Hughes, T. J. R., Linear, Static and Dynamic Finite Element Analysis,
Prentice Hall, 1987.
164