Download as pdf or txt
Download as pdf or txt
You are on page 1of 70

4

THE FINITE ELEMENT METHOD

4-1 INTRODUCTION

The finite element method of analysis is a very powerful, modern computational tool. The method
has been used almost universally to solve very complex structural engineering problems, particularly
in the aircraft industry. The method requires the use of a digital computer because of the large number
of computations involved. Structures take many different, sometimes very diverse, forms and utilize
many different materials. Oftentimes a structure itself is composed of a multitude of materials. In the
design of a bridge, engineers are concerned with the stresses developed in the bridge as a result of the
weight of the vehicles riding on it, in addition to wind loads and its own weight. These stresses are
important because all materials have upper limits on allowable stresses. When these allowable stresses
are exceeded, the structure may no longer perform its intended function. In the extreme case, the
structure may actually collapse [1].

In a typical thermal analysis application, we may need to know the temperatures within the body
being analyzed. Sometimes in addition to the temperatures, we need to know how much heat is flowing
through a given area. In fluid mechanics problems, the analyst may need to determine the fluid
velocities and pressures within or around some device or object. Complexities may arise because of
the irregular and varied geometry, mixed boundary conditions, nonlinear material behavior, and no
uniform loading conditions. The finite element method is particularly well-suited to handle these and
other complications. It should be obvious that it is not possible to obtain closed-form, analytical
solutions to complex problems such as these. Often a skilled engineer can make simplifying
assumptions to obtain a closed-form solution, but the very nature of some problems, may preclude
such an approach.

The basic idea behind the finite element method is to divide the structure, body, or region being
analyzed into a large number of finite elements, or simply elements. These elements may be one, two,
or three dimensional. A popular and classical three-dimensional element is the tetrahedron shown in
Fig. 4-2. When a three-dimensional structure (or heat transfer device, etc.) is divided into hundreds or
sometimes thousands of these no overlapping tetrahedrons, we can see that essentially all volume
geometries can be easily accommodated. Note that this particular element has four nodes (i, j, k, l)
appropriately placed at the vertices of the tetrahedron. Before explaining the purpose of these nodes or
nodal points, let us see how we may use this type of element to represent an irregularly shaped
connecting rod. Fig. 4-1 shows such a connecting rod. By way of illustration let us say that the
temperature distribution within the volume is sought for a given connecting rod material, specified
boundary conditions, etc. The field variable of interest here is the temperature. Because there is an
infinite number of points in the volume, there is also an infinite number of temperatures to be
determined. The problem is therefore said to have an infinite number of degrees of freedom. It should
be recalled that when closed-form analytical solutions exist, they theoretically allow us to compute the
temperature at any point (and hence at an infinite number of them) in the region because the solution
is expressed as a mathematical function of x, y and z [in the form T = f(x,y,z)]. For the connecting rod
in Fig. 4-1, the irregular geometry precludes such a closed form solution. Therefore, an alternate
approach is sought.

Figure 4-1 Connecting rod with variable thickness and an irregular shape.

Instead of requiring that the temperature be determined at every point in the plate, let us change this
unreasonable requirement to a determination of the temperature at only a finite number of points. It
would appear that these points could be taken as the vertices (or nodes) of the tetrahedrons. Figure 4-
2 shows the connecting rod discretized into several tetrahedron elements. Note how the curved
boundaries are approximated by the straight-sided tetrahedrons. The finite element method will
provide the analyst with the temperatures of the connecting rod only at the nodal points. Interpolation
functions are used within each element to describe the variation of the field variable (e.g., here the
field variable is the temperature) as a function of the global coordinates (x, y and z in this case).
Knowledge of the temperatures at the nodal points allows the analyst to ascertain the general
temperature field within the region. These temperatures are often referred to as the nodal temperatures.
From these, the heat flows through each element are also easily determined.

Figure 4-2 Irregularly connecting rod discretized into many tetrahedron finite elements.

As mentioned earlier, a high-speed digital computer is used to perform the large number of
computations involved. An analyst must provide the relevant computer program with the coordinates
of each node, element definitions, material data, boundary condition data, etc. Generally, the accuracy
improves as the number of elements increases, but computational time, and hence cost, also increase.
There is no substitute for experience in trying to determine the number of elements to be used. The
computer program determines the temperature at each node and the heat flows through each element.
The results are usually printed or presented in graphic form.

The analyst may then use this information in the design process. Although the example chosen
above was in the heat transfer area, the same basic idea is applicable to the structural and the fluids
areas, as well as to others. In a structural analysis application, the field variables may be displacements
and/or deflections and slopes. Therefore, in a structural finite element model, we determine the nodal
displacements and/or the nodal deflections and slopes. From these, the stresses and/or bending
moments within an element are easily derived (see Fig. 4-3).

(a) (b)
Figure 4-3 Stressing a plate with a circular hole (a) A square plate with sides length 2a and a circular hole with
radius r under remote tensile loading σ. Misses stress contours calculated with the ABAQUS FE code.

In a fluid flow analysis, the field variables are frequently, but not always, the fluid velocities and
pressure. Therefore, the corresponding finite element model enables a determination of these
parameters at the nodal points. These results may then be used to obtain the shear stresses within each
element if they are needed.

In summarizing, the two key ideas of the finite element method are (1) discretization of the region
being analyzed into finite elements and (2) the use of interpolating polynomials to describe the
variation of a field variable within an element. Both of these notions are developed in subsequent
chapters. The finite element method used today was developed to its present state very recently.
According to Zienkiewicz [2], the development has occurred along two major paths: one in
mathematics and the other in engineering. Somewhere in between these two extremes are the
variational and weighted residual methods, both of which require the use of trial functions to effect a
solution. The use of these functions dates back almost 200 years.

A trial function is an assumed mathematical function that is usually based on physical intuition, is
applied globally to the region being analyzed, and approximates the expected behavior of the region
to some external forces. A qualitative example may help to clarify these abstract points. Let us consider
again the heat conduction problem in a rectangular plate. If the temperature is known to be zero along
the boundaries of the plate and a maximum at the center based on physical reasoning, then one or more
trial functions could be chosen that exhibit this known trend. Such a method is rather limited for two
reasons: (1) the trial functions must satisfy the boundary conditions exactly, and (2) the trial functions
must be chosen based on physical reasoning. In other words, it is very difficult to apply this method to
any but the simplest of problems. The finite element method is predicated along the same lines with
an important exception: the trial functions are not applied globally to the entire region being analyzed
but only locally (i.e., on an element basis). This point is of paramount importance, since the method
becomes applicable to real problems with irregular geometries and unknown solutions. This important
step did not occur until 1943 when Courant [3] introduced piecewise continuous trial functions. As the
name suggests, these functions were not applied globally to the entire region being analyzed but rather
to many small regions or elements. By using trial functions on a local basis, Courant greatly extended
the applicability of trial functions in obtaining approximate solutions to real-world problems. Many
feel that the finite element method has its roots in this paper by Courant, although he did not introduce
the terminology finite element.

4-2 SIMULATIONS USING THE ABAQUS CODE

In a fracture mechanics problem

Fig. 4-4 shows an asymptotic FE mesh using ABAQUS [4] solid CPE4I and cohesive COH2D4
elements, used for a unified treatment of fatigue crack initiation, fatigue crack growth, and stable
tearing crack growth in ductile polycrystalline alloys.

Figure 4-4 Asymptotic FE mesh using ABAQUS solid CPE4I and cohesive COH2D4 elements. Mesh radius
0.1 mm. Shown with the plastic zone around the tip b) FEM mesh showing CTOD profile and damage (in red)
along cohesive path [after ref. 5].

In a welding problem

Table 4-1 shows the welding conditions considered in the simulations. To generate weld and thermal
profiles on thin sheet samples (9 mm x 3 mm x 24 mm) of nickel cast IN-738LC alloy, a 3D transient-
nonlinear finite element model (ABAQUS® code) was developed [6].

Table 4-1: List of GTAW welding conditions used for nickel-based superalloys thin sheet samples
Weld 1 Weld 2 Weld 3
Current (Amps) 150 100 50
Voltage (Volts) 10 10 10
Velocity (mm/s) 5.08 3.387 1.693
Heat Input (J/mm) 296 296 296
Heat distribution parameter c (mm) 4.5 4.5 3

From Fig. 4-5, it is evident the effect of welding parameters on the pool shape and size, and in the
thermal cycles.
(a) (b)
Figure 4-5 Weld profiles and thermal cycles for welds 1, 2 and 3 (a) Top view of weld pool (plane X-Z) and
the solidification temperature range (fusion boundaries and isotherms) (b) Thermal cycles and cooling times
calculated at the documented nodes located at centerline and fusion line.
In Fig. 4-6, residual (Mises) stresses and equivalent plastic strain (PEEQ) contours for Weld 1 after
the steady state reached at 1500 s, are shown.

(a) (b)

(c)
Figure 4-6 Residual (Mises) stresses and equivalent plastic strain (PEEQ) contours for Weld 1 after the steady
state (1500 s) was reached (a) Mises contours in MPa (b) equivalent plastic strain contours (c) equivalent high
elastic strains contours (after ref. 7).

REFERENCES

1. Stassa, F. L., Applied Finite Element Analysis for Engineers. HRW Series in Mechanical
Engineering. CBS Publishing -Japan Ltd. 1985.
2. Zienkiewicz, O. C., The Finite Element Method, McGraw-HiH (UK), London, 1977, pp. 70-72.
3. Courant, R. "Variational Methods for the Solution of Problems of Equilibrium and Vibrations."
Bull. Am. Math. Soc., vol. 49, pp. 1-23. 1943.
4. ABAQUS Finite Element software. SIMULIA Inc.
5. Bonifaz E. A. Cohesive zone modeling to predict failure processes. Canadian Journal on
Mechanical Sciences & Engineering Vol. 2, No. 3, March 2011, pp 42-54.
6. Bonifaz E. A., Richards N. L. Modeling Cast IN-738 Superalloy Gas Tungten-Arc-Welds: Acta
Materialia 57 (2009) 1785-1794.
7. Bonifaz E. A., Richards N. L. Stress-Strain Evolution in Cast IN-738 Superalloy Single Fusion
Welds. International Journal of Applied Mechanics, Vol. 2, No. 4 (2010) 807–826.
5
TRUSS ANALYSIS: THE DIRECT APPROACH

One of the most important practical types of engineering structures is the truss. A truss may be
defined as a structure that is composed of straight members connected at the joints by smooth pins.
Smooth pin joints cannot support a moment. A typical truss is shown in Fig. 5-1. As shown in this
figure, external loads such as P and Q may act only at the joints. A truss must be restrained in some
manner so that it does not undergo free-body motion when loaded. The truss in Fig. 5-1 is restrained
by a pin joint at A and a roller joint at B.

Figure 5-1 A typical two-dimensional truss

Each member in the truss can only support axial loads, which implies that lateral loads may not act
between two joints. All loads, therefore, must act at the joints. Moreover, members in a truss are
connected only at their extremities; hence, members may not be continuous through a joint. In Fig. 5-
1, for example, there are two distinct members, CD and DE, and not one member, CE. A truss is said
to be rigid if it does not collapse when loaded. A rigid truss, however, may and will deform slightly
under the action of one or more loads.

In this chapter, the finite element formulation of a two-dimensional truss is developed [1]. The
direct approach, which does not require the use of variational principles, virtual work principles, or
weighted residual methods, is taken. The main purpose for considering this type of problem is to
demonstrate how the direct approach is used and to provide a step-by-step illustration of the finite
element method. A specific problem is worked out completely by hand, and the basic steps are then
summarized. In section 5-2, a simple ABAQUS model is described and applied to this problem.

5-1 TWO-DIMENSIONAL TRUSS FINITE ELEMENT FORMULATION

Discretization

The first step in any finite element analysis is discretization. Here the truss must be discretized into
a number of finite elements. The truss is composed of axial tensile and compressive members, and it
seems quite natural to consider each member as an element. A typical truss element e is shown in Fig.
5-2. Note that the element has two nodes, i and j, one at either end. It is at these nodes that the forces
are transmitted from one element to the next. These elements are frequently referred to as bar elements,
as opposed to beam or frame elements. Bar elements can withstand only axial forces; beam elements
allow bending moments; and frame elements include axial forces, bending moments, and torsion [4].

Figure 5-2 Typical finite element in truss model.

Each joint in the truss is represented by a node that is given a unique number. Nodes are typically
numbered consecutively from one to the maximum number present. These node numbers are often
referred to as the global node numbers, as opposed to the local node "numbers" i and j. It can be seen
that the way in which the nodes are numbered can have a strong influence on the half-bandwidth (Eq.
5.33), that is, the numbering scheme has a strong influence on the computational time to obtain the
solution. Elements are also numbered consecutively from one to the maximum number of elements.
However, the numbering scheme for the elements is totally arbitrary, unlike the nodal numbering
scheme. Local node numbers in this book are denoted with lowercase letters i, j, k, etc. For example,
element e has only two nodes, and they are denoted as nodes i and j. However, all nodes in a structure
have global node numbers and on element 3, for example, the global node numbers may be 10 (node
i) and 7 (node j). The corresponding global node numbers are denoted with capital letters I, J, K, etc.
The point is that every element in the truss has local node "numbers" i and j with corresponding global
node “numbers” I and J. In the example above, I is 10 and J is 7.

An essential part of every finite element analysis and computer program is the data that defines the
coordinates of every node present in the truss and the elements. Element data are typically composed
of nodal connectivity information and material set flags. Nodal and element data are generally
generated within a computer program, but the user must still provide some minimum amount of data
to describe the particular problem at hand. This is called mesh generation. The ABAQUS models (input
files) presented in Section 5-2, show the data (coordinates, nodes, elements) that defines the problem
domain. Input files have a .inp extension and can be created in any ASCII (text) editor.

Example 5-1

Figure 5-3(a) shows a cantilever truss with only five members and four joints. Members A, C, and
E are composed of a 0.5-in.-diameter steel rod (with a modulus of elasticity of 30 x 106 psi), and
member’s B and D are composed of a 0.4-in. diameter aluminum rod (with a modulus of 11 x 106 psi).
Discretize the truss showing the nodal coordinate data and the element data in tabular form (hence
suitable for use in a computer program). The coordinates (in inches) of the joints are shown in Fig. 5-
3(a).

Figure 5-3 (a) A cantilever truss with five members and four joints. (b) Truss discretized into five elements and
four nodes.
Solution

Since the truss has five members, it is natural to discretize the truss into five elements, each member
being an element. Each joint can then be taken as a node. Therefore, four nodes are used. The first step
in the discretization process is to number each joint or node consecutively from one to the maximum
number of nodes present. For the truss under consideration the nodes are numbered 1 to 4, as shown
in Fig. 5-3(b). Using the coordinates shown in Fig. 5-3 (a) we may summarize the nodal coordinate
data as shown below:

Node number x coordinate, in. y coordinate, in.

1 12.0 0.0
2 12.0 6.0
3 0.0 0.0
4 0.0 10.0

Let us use material set flags 1 and 2 to denote the 0.5-in. steel and 0.4-in. aluminum, respectively.
With the help of Fig. 5-3(b), the element data may be summarized as follows:

Element number Node i Node j Material set flag

1 1 2 1
2 3 1 2
3 3 2 1
4 4 2 2
5 3 4 1

Element Stiffness Relationship in Local Coordinates

It will soon prove to be very convenient if a local coordinate system, x'y', is chosen as shown in Fig.
5-4. This x'y' coordinate system is known as a local coordinate system, since it is defined locally on
the element. Note that the x' axis is directed along the length of the element from node i to node j. The
y' axis must be chosen to be perpendicular to the x' axis, but its sense is arbitrary. Nodal forces in the
x' and y' directions at node i are denoted as Ui'; and Vi', respectively, and nodal displacements as ui'
and vi' respectively. Similarly, at node j, the nodal forces are Uj' and Vj' and the nodal displacements
are uj' and vj' in the x' and y' directions, respectively.

Figure 5-4 Truss element in local coordinate system showing nodal forces and nodal displacements.
It should be recalled from elementary strength of materials that the axial elongation δ is given by
𝑃𝐿
𝛿 = 𝐴𝐸 (5-1)
and,
𝐴𝐸
𝑃= 𝛿
𝐿

where L is the length of the member (or element), P the axial force, A the cross sectional area, and E
the modulus of elasticity (Young's modulus). It is assumed that the elastic range is not exceeded and
that A is constant. Solving Eq. (5-1) for P and interpreting it as the axial force at node i (Ui'), we can
write
𝐴𝐸
𝑈𝑖′ = 𝐿 (𝑢𝑖′ − 𝑢𝑗′ ) (5-2a)

where the effective elongation δ is given by the difference between the nodal displacements along the
element. In a similar fashion, we may write
𝐴𝐸
𝑈𝑗′ = (𝑢𝑗′ − 𝑢𝑖′ ) (5-2b)
𝐿

Note that Ui'= - Uj', which must be the case if element e is to be in static equilibrium in the x' direction.
Because this particular type of element cannot support transverse forces, we also have

𝑉𝑖′ = 0 (5-2c)

𝑉𝑗′ = 0 (5-2d)

Equations 5-2 may be written in matrix form as

1 0 1 0  ui'  U i' 
    
AE  0 0 0 0  vi'  Vi ' 
 (5-3)
L  1 0 1 0 u 'j  U 'j 
    
0 0 0 0 v 'j  V j' 

Since the global nodal coordinates denoted as (xi, yi) and (xj, yj) are specified for any given truss, the
element length may be computed from

𝐿 = √(𝑥𝑗 − 𝑥𝑖 )2 + (𝑦𝑗 − 𝑦𝑖 )2 (5-4)

The other two properties, A and E, are also specified for every element, usually with the help of
material set flags in a computer program. Equation (5-3) may be written concisely as
′ ′ ′
𝐊 𝑒 𝐚𝑒 = 𝐟 𝑒 (5-5)

where Ke' represents the local element stiffness matrix, ae' the local element nodal displacement vector,
and fe' the local element nodal force vector. The superscript (e) is used throughout this text to denote
element and a prime (') to denote the local coordinate system, unless stated otherwise. Before Eq. (5-
5) can be applied to the truss on the whole, it must be transformed to the global (x y) coordinate
system. This transformation is developed in the next section.
Transformation from Local to Global Coordinates

A local (x' y') coordinate system was used to derive the stiffness relationship given by Eq. (5-3).
The x' axis was purposely taken to be directed from node i to node j. The y' axis, which must be chosen
to be perpendicular to the x' axis, is otherwise arbitrary (with respect to its sense). Figure 5-5 shows a
global coordinate system in relation to a typical element e. The purpose of this discussion is to find a
transformation that will enable us to write Eq. (5-5) with respect to a global coordinate system.

Figure 5-5 Relationship between local and global coordinate systems.

First, let us set up a global xy system as shown in Fig. 5-6(a) and draw a position vector r to an
arbitrary point P. The vector r may be written in terms of its two Cartesian components rx and ry as
follows:
𝐫 = 𝑟𝑥 𝐢 + 𝑟𝑦 𝐣 (5-6)

Next, let us define the same vector r in terms of the rotated coordinate system x' y' where, with no
loss in generality, the x' axis is purposely taken to be in the same direction as r itself, as shown in Fig.
5-6(b). In terms of the cartesian components (rx' and ry') in the rotated frame, the vector r becomes

𝐫 = 𝑟𝑥′ 𝐢′ + 𝑟𝑦′ 𝐣′ (5-7)

where i' an j' are the two unit vectors in the x' and y' directions, respectively. From Eqs. (5-6) and (5-
7), we have
𝑟𝑥 𝐢 + 𝑟𝑦 𝐣 = 𝑟𝑥′ 𝐢′ + 𝑟𝑦′ 𝐣′ (5-8)

Figure 5-6 Position vector r (a) in global coordinate system and (b) in local coordinate system.

Taking the dot product of both sides of this equation with respect to i gives

𝑟𝑥 = 𝑟𝑥′ 𝐢. 𝐢′ + 𝑟𝑦′ 𝐢. 𝐣′ (5-9)

where i·i = 1 and i·j = 0 have been used. Note that i·i' represents the cosine of the angle between the
x and x' axes, and i·j' represents the cosine of the angle between the x and y' axes. It should be recalled
that these are the direction cosines, which will be denoted as
𝑛11 = cos(𝑥, 𝑥 ′ ) = 𝐢. 𝐢′ (5-10a)

𝑛12 = cos(𝑥, 𝑦 ′ ) = 𝐢. 𝐣′ (5-10b)

where cos(x,x') represents the cosine of the angle between the x and x' axes, etc. Therefore, with the
help of Eqs. (5-10), Eq. (5-9) becomes
𝑟𝑥 = 𝑛11 𝑟𝑥′ + 𝑛12 𝑟𝑦′ (5-11a)

By forming the dot product of Eq. (5-8) with j in a completely analogous manner we also have

𝑟𝑦 = 𝑛21 𝑟𝑥′ + 𝑛22 𝑟𝑦′ (5-11b)

where n21 represents the cosine of the angle between the y and x' axes, and n22 represents the cosine of
the angle between the y and y' axes. Clearly, in the notation for the direction cosines, the first subscript
is associated with the global coordinate system and the second with the local (rotated) system.
Equations (5-11) may be written in matrix form as
𝑟𝑥 𝑛11 𝑛12 𝑟𝑥′
[𝑟 ] = [𝑛 𝑛22 ] [𝑟𝑦′ ] (5-12a)
𝑦 21
or more concisely as

𝐫 = 𝐓𝐫 ′ (5-12b)

where T is the so-called transformation matrix that transforms a vector in the local system to one in
the global system. It can be shown that the matrix T is orthogonal because T- 1 =TT. Therefore, it
follows from Eq. (5-12b) that
𝐫′ = 𝐓𝑇 𝐫 (5-13)

This last result may be used to transform a vector in the global system to one in a local system. These
transformations are used next to determine the stiffness relationship with respect to the global
coordinate system.

Global Element Stiffness Relationship

Before the element stiffness matrices Ke' and nodal force vectors fe' can be assembled to represent the
entire truss, they must be transformed to a global coordinate system. Interpreting r' as [ui' vi']T and r
as [ui vi]T in Eq. (5-13), we have

𝑢′ 𝑛11 𝑛21 𝑢𝑖
[ 𝑖′ ] = [𝑛 𝑛22 ] [𝑣𝑖 ] (5-14)
𝑣𝑖 12

where ui and vi are the displacements at node i in the global coordinate directions x and y, respectively.
Similarly, for the displacements at node j, it follows from Eq. (5-13) that

𝑢𝑗′ 𝑛11 𝑛21 𝑢𝑗


[ ′ ] = [𝑛 𝑛22 ] [𝑣𝑗 ] (5-15)
𝑣𝑗 12

Figure 5-7(a) shows the relationship between the local and global nodal displacements. Equations (5-
14) and (5-15) may be written in one matrix equation as
Figure 5-7 Typical element e showing global and local (a) nodal displacements and (b) nodal forces. All forces
and displacements are shown with a positive sense.

 ui'  n11 n21 0 0   ui 


 '   
 vi   n12 n22 0 0   vi 
(5-16)
u 'j   0 0 n11 n21  u j 
 '   
v j   0 0 n12 n22  v j 

or more concisely as

𝐚𝑒 = 𝐑𝐚𝑒 (5-17)

where, by definition

TT 0
R  (5-18)
0 TT 

and ae represents the vector of nodal displacements for element e referred to the global coordinate
system, or


ae  ui vi uj vj 
T
(5-19)

In a completely analogous manner we may also write



𝐟 𝑒 = 𝐑𝐟 𝑒 (5-20)

where fe is the vector of nodal forces for element e referred to the global coordinate system, or


f e  U i Vi U j V j  T
(5-21)

Figure 5-7(b) shows the relationship between the local and global nodal forces. With the help of
Eqs. (5-17) and (5-20), we may write Eq. (5-5) as

𝐊 𝑒 𝐑𝐚𝑒 = 𝐑𝐟 𝑒 (5-22)

which may then be premultiplied by RT to give



𝐑𝑇 𝐊 𝑒 𝐑𝐚𝑒 = 𝐟 𝑒 (5-23)
where RTR = R- 1 R = I have been used on the right-hand side. Equation (5-23) may be written simply
as
𝐊 𝑒 𝐚𝑒 = 𝐟 𝑒 (5-24)

where Ke, the element stiffness matrix referred to global coordinates, is obtained directly from the
element stiffness matrix referred to local coordinates (Ke') by

𝐊 𝑒 = 𝐑𝑻 𝐊 𝒆 𝐑 (5-25)

The matrix Ke is often simply referred to as the global element stiffness matrix (as opposed to the
local element stiffness matrix Ke'). If the multiplications indicated in Eq. (5-25) are carried out, we get

 n112 n21n11  n112  n21n11 


 2 
AE  n11n21 2
n21  n11n21  n21 
K 
e

L  n11  n21n11
2 2
n11 n21n11 
(5-26)
 
 n11n21  n21 
2 2
n11n21 n21

It should be noted that in the expression for Ke only two of the four possible direction cosines
appear: n11 and n21. By definition, n11 is the cosine of the angle between the x and x' axes, and n21 is
the cosine of the angle between the y and x' axes. Recall that the x' axis is directed along the element
from node i to node j. From Fig. 5-8 it is immediately apparent that

Figure 5-8 Definition of angles θx and θ y used in obtaining the direction Cosines.
𝑥𝑗 −𝑥𝑖
𝑛11 = cos 𝜃𝑥 = (5-27-a)
𝐿

and
𝑦𝑗 −𝑦𝑖
𝑛21 = cos 𝜃𝑦 = (5-27b)
𝐿

where (xi, yi) and (xj, yj) for element e are the global coordinates of nodes i and j, respectively, and L
is the element length [easily computed from Eq. (5-4)]. It should be recalled that the coordinates of
each node are a vital part of the input to every finite element analysis (and computer program).
Consequently, the two needed direction cosines are routinely computed with Eqs. (5-27). It must be
emphasized that the important result is Eq. (5-26), which enables us to compute the global element
stiffness matrix for every element since A and E are specified, and L, n11, and n21 are easily computed
(for every element). In the next section, the notion of assemblage is introduced. It is an easy matter to
show that Ke is symmetric because Ke' is itself symmetric.
Example 5-2

For the truss in Example 5-1, determine the global element stiffness matrix for element 3.

Solution

The global element stiffness matrix, denoted as Ke, is given by Eq. (5-26) in terms of the direction
cosines n11 and n21, which are computed from Eqs. (5-27) for element 3 as follows:

𝑥2 − 𝑥3 12.0 − 0.0
𝑛11 = = = 0.8944
𝐿(3) 13.42

𝑦2 − 𝑦3 6.0 − 0.0
𝑛21 = = = 0.4472
𝐿(3) 13.42

Here, Eq. (5-4) has been used to get the element length L (3). The calculation of K (3)
from Eq. (5-26)
is now straightforward, and the result is

 351 176  351  176


 88  176  88 
3  176
K ( 3)  10 lbf/in .
 351  176 351 176 
 
 176  88 176 88 

where the cross-sectional area A(3) was calculated for the 0.5-in.-diameter steel rod (with a modulus of
elasticity of 30 x 106 psi). As expected, K (3) is symmetric.

Assemblage

The assemblage step is necessary in all finite element analyses. It is in this step that the original
truss (or in general, any structure, body, or region being analyzed) is put back together or assembled
from the individual elements comprising it. Suffice it to say now that the assemblage step is based on
the principle of compatibility; that is, the x and y displacements associated with a particular node on
any given element must be identical to those associated with the same node on each element that shares
this node [4].

Consider a two-dimensional truss with N nodes. Each element e has two nodes i and j with global
node numbers I and J. Since each node has two degrees of freedom (i.e., two components of
displacement) the 4 x 4 global element stiffness matrix may be thought of as being partitioned into
four submatrices, each of size 2 x 2, or

 K iie K ije 
K  e
e
e  (5-28)
K ji K jj 

These 2 x 2 submatrices are easily extracted from Eq. (3-26). Since global node number I corresponds
to local node i and global node number J to local node j, Eq. (5-28) may be written in terms of the
global nodal subscripts as
K eI ,I K eI ,J 
K  e
e
 (5-29)
K J ,I K eJ ,J 

The assemblage stiffness matrix, denoted as Ka, is always of the form


𝑎 𝑎 𝑎 𝑎
𝐊1,1 𝐊1,2 𝐊1,3 ⋯ 𝐊1,𝑁
𝐊 𝑎2,1 𝐊 𝑎2,2 𝐊 𝑎2,3 ⋯ 𝐊 𝑎2,𝑁
𝐊 𝑎 = 𝐊 𝑎3,1 𝐊 𝑎3,2 𝐊 𝑎3,3 ⋯ 𝐊 𝑎3,𝑁 (5-30)
⋮ ⋮ ⋮ ⋱ ⋮
𝐊 𝑎 𝐊 𝑎𝑁,2 𝐊 𝑎𝑁,3 ⋯ 𝐊 𝑎𝑁,𝑁 ]
[ 𝑁,1

Note that the superscript (a) denotes assemblage. Each of these KaI,J 's are 2 x 2 submatrices. To begin
the assemblage process, the assemblage stiffness matrix Ka is zeroed out. Then the individual global
element stiffness submatrices (i.e., the KeI,J 's) are added to the corresponding position in the matrix
Kª (i.e., at KaI,J) for element 1, then for element 2, and so forth, up to and including the last element.
The assemblage of each 2 x 2 submatrix in Eq. (5-30) may be given by the relationship

M
K aI ,J   K eI ,J for I  1 to N and J  1 to N (5-31)
e1

In Eq. (5-31), M is the maximum number of elements, and KeI,J is taken as the 2 x 2 null matrix if
element e does not have nodes I and J. The next example will help to clarify the assemblage process
for the truss.

Example 5-3

For the truss of Example 5-1, determine the assemblage stiffness matrix in terms of the 2 x 2 global
element stiffness submatrices KeI,J.

Solution

Element 1 has global node numbers 1 and 2 (i.e., I and J, respectively), and so the global element
stiffness matrix for element 1, denoted as K (1), is given by

K1(1,1) K1(1, 2) 
K (1)
  (1) (1) 
K 2,1 K 2, 2 

Having first zeroed-out the assemblage stiffness matrix Ka, we may then “add'' this result to Ka to get
K1(1,1) K1(1, 2) 0 0
 (1) 
 K 2,1 K (21,)2 0 0
K 
a
 0 0 0 0
 
 0 0 0 0

This is not the complete form of the assemblage stiffness matrix because four more elements have yet
to be considered. Element 2 has global node numbers 3 and l (i.e., I and J, respectively) and the global
element stiffness matrix for element 2 is given by

K 3( 2,3) K 3( 2,1) 
K ( 2)
  ( 2) 
K1,3 K1(,21) 

After combining this element stiffness matrix with the result from element 1, we get

K1(1,1)  K1(,21) K1(1, 2) K1(,23) 0


 
K (21,1) K (21,)2 0 0
K 
a
 K 3,1 ( 2 )
0 K 3( 2,3) 0
 
 0 0 0 0

Note the position of each entry in Kª carefully. After considering element 3 (with nodes 3 and 2),
element 4 (with nodes 4 and 2), and element 5 (with nodes 3 and 4), the assemblage stiffness matrix
becomes
K1(1,1)  K1(,21) K1(1, 2) K1(,23) 0 
 

(1)
K 2,1 K 2, 2  K 2, 2  K 2, 2
(1) ( 3) ( 4) ( 3)
K 2, 3 ( 4)
K 2, 4 
K 
a
 K 3( 2,1) K 3(3, 2) K 3( 2,3)  K 3(3,3)  K 3(5,3) K 3(5, 4) 
 
 0 K (44, 2) K (45,3) K (44, 4)  K (45, 4) 
Several significant observations should be made. First, each of the individual 2 x 2 element stiffness
submatrices is symmetric, as an examination of Eq. (5-26) clearly shows. Mathematically, this may be
stated as
𝐊 𝑒𝐼,𝐽 = 𝐊𝐽,𝐼
𝑒
(5-32)

It then follows that Kª itself is symmetric, which has several advantages as delineated in Chapters 6
and 7. Another property of Kª is that it is banded, which means the nonzero entries in Kª are gathered
along the main diagonal. In the example above, this is not very obvious because the truss has only five
elements and four nodes. Because of the symmetric nature of Kª, we usually use the term
half-bandwidth, denoted as bw in this text. A general expression for the half-bandwidth is given by

𝑏𝑤 = 𝑁DOF (1 + ∆max ) (5-33)

where NDOF is the number of degrees of freedom per node and Δ max is the maximum of the maximum
differences between node numbers on each element considering all elements one at a time. In the truss
model there are two degrees of freedom associated with each node (the two components of
displacement). From Eq. (5-33) it can be seen that the way in which the nodes are numbered can have
a strong influence on the half-bandwidth. In a computer program, the execution time increases with
the square of the bandwidth for the usual solution techniques [4].

Example 5-4

Determine the half-bandwidth of the assemblage stiffness matrix from Example 5-3 by direct
examination of Kª and by computation using Eq. (5-33).

Solution

Let us write Kª in the following form showing each potentially nonzero entry as an x and each zero
entry as a 0:

A direct count of the number of nonzero entries in the half-bandwidth above yields 6. Equation (5-
33), of course, must give the same result for the half-bandwidth, but let us compute it. There are two
degrees of freedom per node, hence NDOF = 2. The maximum difference between any two node
numbers considering all elements one at a time is 2, or Δmax = 2. Therefore, Eq. (5-33) readily gives
bw= 2(l +2) = 6 as expected.

Application of Loads

The assemblage nodal force vector fª may be obtained by an assemblage process similar to the one
used to obtain Ka. It is simpler, however, because each element nodal force vector, fe, is a column
vector (of size 4 x 1), as opposed to a matrix. The assemblage of each element nodal force vector in
this manner is illustrated in subsequent sections. Because the external loads on the truss can only occur
at the joints, the effective assemblage nodal force vector fª can be obtained by a more direct route as
explained below.

The assemblage nodal force vector, with 2N entries, is first zeroed-out, where N is the number of
nodes present. Let us assume that at global node number K, there is a point load with components Fx
and Fy. These components are referred to the global coordinate system as shown in Fig. 5-9. Because
the load is applied at node K, we must add Fx and Fy to the (2K-1)st and 2Kth positions in fa,
respectively. Reaction forces are not considered to be loads in finite element analysis. They are handled
routinely as restraints.
Figure 5-9 Portion of truss showing node K with external load (components Fx and Fy)

Example 5-5

For the loading shown in Fig. 5-3(a) for the truss in Example 5-1, determine the assemblage nodal
force vector fª.

Solution

The 2000-lbf load is the only external load and must be resolved into its two Cartesian components,
Fx and Fy. Note from Example 5-1 that an xy (global) coordinate system is implied as shown in Fig. 5-
3(a). The x and y components of the applied load are given by Fx = - 2000 cos 60º = - l 000 lb and Fy=
- 2000 sin 60º = - 1732 lb. Since this load is applied to node 1 (node K), the first (from 2K–1) and
second (from 2K) entries in fª must be changed to yield the following result:

 1000 Fx 
 1732 Node1
Fy 
 
 0  Fx 
  Node2
 0  Fy 
f 
a
 0  Fx 
  Node3
 0  Fy 
 0  Fx 
  Node4
 0  Fy 

If another force also acts at node 1, the respective components would be added to those above. The
restraints on the nodal displacements are considered next.

Application of Restraints on Nodal Displacements and Solution

The assemblage stiffness matrix Kª and nodal force vector fª are obtained as described above. The
assemblage system equation is of the form

𝐊𝑎 𝐚 = 𝐟𝑎 (5-34)

where a is the vector of nodal unknowns or, in this case, the vector of unknown nodal displacements.
The vector a is given by
a  u1 v1 u2 v2 ... uN vN 
T
(5-35)

where N is the maximum number of nodes and uI, vI are the x and y components of the displacement
at global node number I. In Eq. (5-35), the transpose is used to conserve space - a itself is a column
vector as implied by Eq. (5-34). Equation (5-34) cannot be solved for the vector a because the restraints
on the nodal displacements have not been considered. In fact, the inverse of Kª does not exist, in
general, until the restraints on displacements are taken into account. In other words, Kª is singular.
After the restraints are considered, Eq. (5-34) is written as

𝐊𝐚 = 𝐟

where K is no longer singular. Therefore, K-1 exists, which implies a unique solution for the vector of
nodal displacements a. Two methods are commonly used to impose the prescribed displacement
restraints [4]. Each is discussed below.

Method 1

The first method is most easily explained by an example. Consider the system of equations

𝑘11 𝑎1 + 𝑘12 𝑎2 + 𝑘13 𝑎3 = 𝑓1 (5-36a)


𝑘21 𝑎1 + 𝑘22 𝑎2 + 𝑘23 𝑎3 = 𝑓2 (5-36b)
𝑘31 𝑎1 + 𝑘32 𝑎2 + 𝑘33 𝑎3 = 𝑓3 (5-36c)

Let us impose a restraint on a2, in other words, that a2 is to have a specified value, e.g., ap (a prescribed
displacement in this truss analysis). Let us further assume that the system of equations is symmetric
(such as the truss model) and that the symmetry is to be preserved. Let us replace Eq. (3-36b) with

𝑎 2 = 𝑎𝑝

The system of equations thus becomes

𝑘11 𝑎1 + 𝑘12 𝑎2 + 𝑘13 𝑎3 = 𝑓1


𝑎2 = 𝑎𝑝
𝑘31 𝑎1 + 𝑘32 𝑎2 + 𝑘33 𝑎3 = 𝑓3

But the symmetry has been destroyed, and to preserve it we need a zero coefficient on a2 in the first
equation. We can get that by noting that a2 = ap (which is known) and writing

𝑘11 𝑎1 + 𝑘13 𝑎3 = 𝑓1 − 𝑘12 𝑎𝑝


Note that the term involving a2 is conveniently transposed to the right-hand side (and ap replaces a2).
After performing a similar operation on the third equation and using matrix notation, we get

𝑘11 0 𝑘13 𝑎1 = 𝑓1 − 𝑘12 𝑎𝑝


[ 0 1 0 ] [𝑎 2 ] = [ 𝑎𝑝 ] (5-37)
𝑘31 0 𝑘33 𝑎3 = 𝑓3 − 𝑘32 𝑎𝑝

Note that f2 disappears! The implication is that for degrees of freedom in the truss model where an
unknown reaction force exists, this unknown force (f2 above) never really enters the formulation.
Instead, a known displacement is imposed. In structural analysis problems, these prescribed
displacements are usually zero, and in this case ap = 0 in Eq. (5-37).

Method 2

A much simpler method exists for imposing prescribed displacement restraints. The method is based
on the concept of penalty functions, and a proof of why the method works is given in Zienkiewicz [
1]. The purpose here is not to prove it but to learn how to apply it. Let us again consider the system of
equations in Eq. (5-36). In this method a large number β is selected, say six to twelve orders of
magnitude larger than the largest coefficient kij, and is added to the coefficient kii if ai is to be
prescribed. In addition, the right-hand side of the ith equation is changed to β times the prescribed
value (of displacement). In the example above, since a2 is to be prescribed (i.e., a2 = ap), the second
equation becomes

𝑘21 𝑎1 + (𝑘22 + β)𝑎2 + 𝑘23 𝑎3 = β𝑎𝑝

Note that β is chosen to be much larger than the coefficients and, for all practical purposes, this last
equation is really equivalent to
β𝑎2 = 𝛽𝑎𝑝
or
𝑎 2 = 𝑎𝑝

which is the desired result. The other equations are unchanged. In matrix form the system of equations
becomes
𝑘11 𝑘12 𝑘13 𝑎1 𝑓1
[𝑘21 𝑘22 + β 𝑘23 ] [𝑎2 ] = [β𝑎𝑝 ] (5-38)
𝑘31 𝑘32 𝑘33 𝑎 3 𝑓3

Note that this simple method automatically preserves symmetry. In a computer program, Method 2 is
extremely easy to apply. Both methods result in the system of equations represented by the matrix
equation (without any superscripts).
𝐊𝐚 = 𝐟 (5-39)

This equation may be solved for the nodal displacements contained in the vector a by any method
applicable to the solution of linear, algebraic equations. Here, the matrix inversion method will be
used. If K-1 is computed and if Eq. (5-39) is premultiplied by it, we get

𝐚 = 𝐊 −1 𝐟 (5-40)

This matrix inversion method is used in the solution of example problems.

Computation of the Element Resultants

By definition, the element resultants for the truss include quantities such as the axial elongations,
strains, stresses, and forces. Each of these is computed more or less directly from the nodal
displacements a that were computed to be the solution of Ka = f.

The elongation δ for element e may be computed in terms of the x' components of nodal
displacements at nodes i and j by
𝛿𝑗−𝑖 = 𝑢𝑗′ − 𝑢𝑖′ (5-41a)
𝛿𝑖−𝑗 = 𝑢𝑖′ − 𝑢𝑗′ (5-41b)

But from Eq. (5-14), we have


𝑢𝑖′ = 𝑛11 𝑢𝑖 + 𝑛21 𝑣𝑖 (5-42a)

and from Eq. (5-15),


𝑢𝑗′ = 𝑛11 𝑢𝑗 + 𝑛21 𝑣𝑗 (5-42b)

Note that in Eqs. (5-42) only two of the direction cosines appear (n11 and n21): these are easily
computed for every element by Eqs. (5-27). In Eqs. (5-42), ui, vi, uj, and vj are known once the nodal
displacements are determined (from the solution of Ka = f). Therefore, for any element e, the
elongation δ is readily computed. The axial strain ɛ for a typical element e is computed from

δ
ε=𝐿 (5-43)

where L is the element length computed from Eq. (5-4). The material is assumed to be elastic so that
Hooke's law applies. Therefore, the axial stress σ in the element is computed from

σ = 𝐸ε (5-44)

Finally, the axial force F in the element is determined by

𝐹 = σ𝐴 (5-45)

In Eqs. (5-41), (5-43), (5-44), and (5-45), positive values of δ, ɛ, σ, and F denote tension, whereas
negative values denote compression. The formulation of this section is utilized in the next section in
an example.

Application to Example 5-1

In this section, a simple problem is executed completely by hand in order to illustrate the basic steps
in the finite element method. Let us reconsider the truss in Example 5-l shown in Fig. 5-3(a). It should
be recalled that members A, C. and E are made of a 0.5-in.-diameter steel rod and members B and D
of a 0.4-in.-diameter aluminum rod. The modulus of elasticity of steel and aluminum are 30 x l 0 6 and
11 x 106 psi, respectively. The truss is constrained and loaded as shown in Fig. 5-3(a), which also
shows the coordinates of each joint.
The first step in any finite element analysis is discretization. Let us use the results of Example 5- l
where we took each member of the truss to be an element. It should be recalled from that example that
the discretization step may be divided into two separate but related tasks: (1) specification of the nodal
coordinates and (2) specification of the element data, including material set definitions. Because this
step is so vitally important, let us summarize the results of the discretization from Example 5-1:

Nodal Coordinate Data

Node number x coordinate, in. y coordinate, in.

1 12.0 0.0
2 12.0 6.0
3 0.0 0.0
4 0.0 10.0

Element Data

Element number Node i Node j Material set

1 1 2 1
2 3 1 2
3 3 2 1
4 4 2 2
5 3 4 1

Note that member A is taken as element 5, member B as element 4, and so forth. In addition, material
set 1 is the 0.5-in. steel and material set 2, the 0.4-in. aluminum. The nodal numbering scheme used
above will result in the smallest possible bandwidth of the assemblage stiffness matrix.

In Eq. (5-26), three properties are needed for each element: the cross-sectional area, the modulus of
elasticity, and the length. The length is easily calculated by Eq. (5-4) from the nodal coordinates.
However, the remaining two ''material properties'' must be specified:

Material set Cross-sectional area, in.2 Modulus of elasticity, psi

1 0.1963 30 x 10 6
2 0.1257 11 x 10 6

All of the above information is used frequently throughout the finite element solution process. In
summary, the truss model will be composed of four nodes and five elements involving two different
types of materials. Not surprisingly, this same information becomes an integral part of the ABAQUS
input files described in Sec. 5-2.

The next step is to determine the local element stiffness matrix for element 1, to transform it to the
global element stiffness matrix, and finally to ''add'' it into the assemblage stiffness matrix (which is
always zeroed-out before processing the first element). With the help of Eq. (5-26)
 n112 n21n11  n112  n21n11 
 2 
AE  n11n21 2
n21  n11n21  n21 
K 
e

L   n112  n21n11 2
n11 n21n11 
(5-26)
 
 n11n21  n21 
2 2
n11n21 n21

the global element stiffness matrix for element l may be verified to be

0 0 0 0
6  0  1
(0.1963)(30x10 ) 0 1
K (1)  lbf/in.
(12.  12.) 2  (6.  0.) 2 0 0 0 0
 
0  1 0 1

0 0 0 0 
0 982 0  982
K (1)  103  lbf/in.
0 0 0 0 
 
0  982 0 982 

Recall that element 1 is composed of material 1 and that only two of the direction cosines are needed,
which, for element 1, are readily computed as follows:

𝑥𝑗 − 𝑥𝑖 12.0 − 12.0
𝑛11 = = = 0.0
𝐿 6.0
𝑦𝑗 − 𝑦𝑖 6.0 − 0.0
𝑛21 = = = 1.0
𝐿 6.0

The reader should examine the orientation of element 1 in Fig. 5-4(b) to verify these direction cosines
by inspection. This result must now be added into the assemblage stiffness matrix (which at this point
is simply a null matrix). Since element l connects nodes l and 2, the assemblage stiffness matrix Kª
becomes
0 0 0 0 0 0 0 0
0 982 0  982 0 0 0 0

0 0 0 0 0 0 0 0
 
0  982 0 982 0 0 0 0
K a  103  lbf/in .
0 0 0 0 0 0 0 0
 
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
 
0 0 0 0 0 0 0 0

The reader should corroborate this result for the assemblage stiffness matrix (after processing the first
element) with Example 5-3. Although the nodal force vectors fe, and fa could be considered at this
point for element 1, it proves to be more convenient to obtain fa directly as indicated by Example 5-5.
This same comment applies, of course, to the nodal force vectors for elements 2 through 5.
Although only three or four significant digits are shown here, all calculations were carried out using
nine digits and therefore, numbers may not add exactly as shown during the assemblage. The reader
should note the scalar multiplier of 103 and verify that each entry in K(1) has units of pound-force per
inch. The stiffness terms in structural models will always have units of force per unit length-similar
to the units of the stiffness for a spring.

Element 2 is considered next, and since it is composed of material set 2, K(2) is given by

1 0 1 0
 6
0
(0.1257)(11x10 )  0 1 0
K ( 2)  lbf/in.
12.0  1 0 1 0
 
0 0 0 0

 115 0  115 0
 0 1 0 0
K ( 2)  103  lbf/in.
 115 0 115 0
 
 0 0 0 0

The reader may verify that for element 2 we have n11 = l .0 and n21 = 0.0, both by inspection and by
using Eqs. (5-27). Therefore, we have K(2) = K(2)', or the global element stiffness matrix is identical to
the local element stiffness matrix [see Eq. (5-26)]. Since element 2 is connected by nodes 3 and 1, the
assemblage stiffness matrix becomes (after processing two elements)

 115 0 0 0  115 0 0 0
 0 982 0  982 0 0 0 0

 0 0 0 0 0 0 0 0
 
0  982 0 982 0 0 0 0
K a  103  lbf/in .
 115 0 0 0 115 0 0 0
 
 0 0 0 0 0 0 0 0
 0 0 0 0 0 0 0 0
 
 0 0 0 0 0 0 0 0

Element 3 is a little more interesting than the previous two elements because it is oriented neither
horizontally nor vertically. The length of element 3 is easily computed from Eq. (5-4) to be

𝐿 = √(12. −0−)2 + (6. −0. )2 = 13.42 𝑖𝑛.

and K(3) becomes


 n112 n21n11  n112  n21n11 
 2 
(0.1963)(30x106 )  n11n21 2
n21  n11n21  n21 lbf/in.
K (3) 
13.42   n11  n21n11
2 2
n11 n21n11 
 
 n11n21  n21 
2 2
n11n21 n21

Recall from Example 5-2, for element 3 we calculated n11 = 0.894, n21 = 0.447. The global element
stiffness matrix for element 3 is computed to be
 351 176  351  176
 176 88  176  88 
K ( 3)  103  lbf/in .
 351  176 351 176 
 
 176  88 176 88 

This matrix must now be included in the assemblage matrix Ka, and since element 3 connects nodes 3
and 2, we get the result

 115 0 0 0  115 0 0 0
 0 982 0  982 0 0 0 0

 0 0 351 176  351  176 0 0
 
0  982 176 1070  176  88 0 0
K a  103  lbf/in .
 115 0  351  176 466 176 0 0
 
 0 0  176  88 176 88 0 0
 0 0 0 0 0 0 0 0
 
 0 0 0 0 0 0 0 0

Once again the reader should reexamine Example 5-3. This last result is the assemblage stiffness matrix
after processing three elements. The procedure to be used on element 4 is identical to that used above.
The reader should verify that for element 4 with n11 = 0.949 and n21 = -0.316, we get

 98  33  98 33 
 33 11 33  11
K ( 4)  103  lbf/in .
 98 33 98  33
 
 33  11  33 11 

After processing four elements and noting that element 4 connects nodes 4 and 2, the assemblage
matrix Kª becomes

 115 0 0 0  115 0 0 0 
 0 982 0  982 0 0 0 0 

 0 0 450 143  351  176  98 33 
 
0  982 143 1080  176  88 33  11
K a  103  lbf/in .
 115 0  351  176 466 176 0 0 
 
 0 0  176  88 176 88 0 0 
 0 0  98 33 0 0 98  33
 
 0 0 33  11 0 0  33 11 

The pertinent results for element 5 are now summarized. The global element stiffness matrix K(5) with
n11 = 0.0 and n21 = 1.0 becomes

0 0 0 0 
0 589 0  589
K ( 5)  103  lbf/in .
0 0 0 0 
 
0  589 0 589 
Since element 5 connects nodes 3 and 4 (in this order), the assemblage stiffness matrix becomes

 115 0 0 0  115 0 0 0 
 0 982 0  982 0 0 0 0 

 0 0 450 143  351  176  98 33 
 
3 0  982 143 1080  176  88 33  11  (5-46)
K  10
a
lbf/in .
 115 0  351  176 466 176 0 0 
 
 0 0  176  88 176 677 0  589
 0 0  98 33 0 0 98  33 
 
 0 0 33  11 0  589  33 600 

All five elements have now been processed, and this last version of the assemblage stiffness matrix
represents the complete assemblage stiffness matrix for the entire truss. Note that it is symmetric and
slightly banded with a half-bandwidth of 6 as observed from the third row (also see Example 5-4).

The next step is to determine the assemblage nodal force vector. Because the truss and loading in
the present example are the same as those in Example 5-5, we may use the result of that example
directly. For completeness, fª is restated here:

𝐟 𝑎 = [−1000 −1732 ⋮ 0 0 ⋮ 0 0 ⋮ 0 0]𝑇 (5-47)

Note the use of the transpose to denote fa, a column vector in row matrix form.

Actually, there are three other loads as a result of the reactions. As mentioned above, these reaction
forces need not be considered in the construction of fª. However, let us include these reactions in fª to
review why this is the case. Let us denote the reaction in the x direction at node 3 as U3 and denote the
reactions in the x and y directions at node 4 as U4 and V4, respectively. If these external forces are
included in the assemblage nodal force vector, we get

𝐟 𝑎 = [−1000 −1732 ⋮ 0 0 ⋮ 𝑈3 0 ⋮ 𝑈4 𝑉4 ]𝑇 (5-48)

The assemblage system equation Kªa = fa becomes

 115 0 0 0  115 0 0 0   u1   1000


 0
 982 0  982 0 0 0 0   v1   1732
 0 0 450 143  351  176  98 33  u 2   0 
    
3 0  982 143 1080  176  88 33  11  v2   0  (5-49)
10 
 115 0  351  176 466 176 0 0  u3   U 3 
    
 0 0  176  88 176 677 0  589  v3   0 
 0 0  98 33 0 0 98  33  u 4   U 4 
    
 0 0 33  11 0  589  33 600  v4   V4 

At this point, the matrix on the left is singular; that is, its inverse does not exist and, therefore, it is
impossible to so1ve for a, the vector of unknown nodal displacements. However, the restraints on
some of the nodal displacements must be considered, namely u3 = 0, u4 = 0, and v4 = 0. Let us use the
second method (Method 2) to impose these restraints. (The first method is illustrated numerically in
Chapter 6.) An arbitrary value of 1.0 x l018 is taken as β, and Eq. (5-49) becomes
 115 0 0 0  115 0 0 0   u1    1000 
 0
 982 0  982 0 0 0 0   v1    1732 
 0 0 450 143  351  176  98 33  u 2   0 
    
3 0  982 143 1080  176  88 33  11  v2   0 
10  (5-50)
 115 0  351  176 1015 176 0 0  u3  0 x 1018 
    
 0 0  176  88 176 677 0  589  v3   0 
 0 0  98 33 0 0 1015  33  u 4  0 x 1018 
    
 0 0 33  11 0  589  33 1015  v4  0 x 1018 

The 1015 (and not 1018) is entered on the appropriate diagonal entries because of the l03 scalar
multiplier. Note that after the restraints are considered, the right-hand side is effectively

𝐟 = [−1000 −1732 ⋮ 0 0 ⋮ 0 0 ⋮ 0 0] 𝑇

which is precisely the same as Eq. (5-47). Therefore, the loads as a result of the reactions need not be
considered a very fortuitous result.

If Eq. (5-50) is solved by the matrix inversion method, the resulting nodal displacements, in inches,
are
𝑢1 = −0.00868
𝑣1 = −0.03528
𝑢2 = +0.00996
𝑣2 = −0.03351
𝑢3 = 0.00000
𝑣3 = −0.00176
𝑢4 = 0.00000
𝑣4 = 0.00000

Note that the restrained degrees of freedom all have zero displacements, as expected.

The only remaining task is to determine the so-called element resultants, which include the axial
elongations, strains, stresses, and forces. Element 3 is considered in detail to illustrate the use of Eqs.
(5-41) to (5-45). The final results for all elements are subsequently summarized below. Recall that
element 3 connects nodes 3 and 2 (in this order). From Eqs. (5-41) and (5-42) with n11 = 0.894 and
n21=0.447, the axial elongation δ for element 3 becomes

𝛿𝑗−𝑖 = 𝑢𝑗′ − 𝑢𝑖′

𝛿𝑗−𝑖 = (𝑛11 𝑢𝑗 + 𝑛21 𝑣𝑗 ) − (𝑛11 𝑢𝑖 + 𝑛21 𝑣𝑖 )

(3)
𝛿𝑗−𝑖 = [(0.894)(0.00996) + (0.447)(−0.03351)] − [(0.894)(0.0) + (0.447)(−0.00176)]

(3)
𝛿𝑗−𝑖 = −0.00529 in.
i j

or
𝛿𝑖−𝑗 = 𝑢𝑖′ − 𝑢𝑗′

𝛿𝑖−𝑗 = (𝑛11 𝑢𝑖 + 𝑛21 𝑣𝑖 ) − (𝑛11 𝑢𝑗 + 𝑛21 𝑣𝑗 )

(3)
𝛿𝑖−𝑗 = [(0.894)(0. ) + (0.447)(−0.00176)] − [(0.894)(0.00996) + (0.447)(−0.03351)]

(3)
𝛿𝑖−𝑗 = 0.00529 in. as expected
i j

Note that the bar is in compression. The strain in element 3 is readily computed by Eq. (5-43) as

(3)
δ(3) −0.00529 𝑖𝑛.
ε = (3) = = −394 𝑥 10−6
𝐿 13.42 𝑖𝑛.

the stress from Eq. (5-44) as

σ(3) = ε𝐸 = (−394 x 10−6 )(30 x 106 ) = −11,800 psi

and finally the force as

𝐹 (3) = σ𝐴 = (−11,800)(0.1963) = 2320 lbf compression

The reader should determine the axial elongations, strains, stresses, and forces for the remaining
elements. The results are summarized in Table 5-1 below.

Table 5-1 Results for Example 5-1

Element number δ, in ɛ, in./in. σ, psi F, lbf

1 0.00176 0.000294 8,820 1730


2 -0.00868 -0.000723 -7,960 -1000
3 -0.00529 -0.000395 -11,800 -2320
4 0.02004 0.001585 17,400 2190
5 0.00176 0.000176 5,290 1040

Note that elements l, 4, and 5 (members E, B, and A) are in tension and elements 2 and 3 (member’s
D and C) are in compression. The maximum force of 2320 (compression) occurs in element 3 (member
C) but the maximum stress of 17,400 psi (tension) occurs in element 4 (member B). The reader is asked
to determine the reaction forces U3, U4, and V4 directly with the help of Eq. (5-49).

This completes the two-dimensional truss example. In Sec. 5-2, several ABAQUS models (input
files) illustrate how all of the steps executed above can be automated.

5.2 STRUCTURAL EXAMPLES USING THE ABAQUS CODE

Example 1: Two dimensional cantilever truss for Example 5-1

Figure 5-3 (described in section 5-1) shows a cantilever truss consisting of five members and four
joints. Members A, C, and E are composed of a 0.5-in.-diameter steel rod (with a modulus of elasticity
of 30 x 106 psi), and member’s B and D are composed of a 0.4-in. diameter aluminum rod (with a
modulus of 11 x 106 psi). Develop an ABAQUS suitable input file to print the stress (S11) and strain
(E11) for all truss members and compare with the results shown on Table 5-1. The coordinates (in
inches) of the joints are shown in Fig. 5-3(a).

Figure 5-3 (a) A cantilever truss with five members and four joints. (b) Truss discretized into five
elements and four nodes.

Input file: barras.inp

*HEADING **
Two Dimensional Truss * STEP, PERTURBATION
**Written by E. A. Bonifaz 2000 lb inclined load
*PREPRINT, ECHO=YES, MODEL=YES, HISTORY=YES *STATIC
** *BOUNDARY
*NODE, NSET=TODOS 4, 1, 2
1, 12., 0., 0. 3, 1
2, 12., 6., 0. * CLOAD
3, 0., 0., 0. 1, 2, -1732.05
4, 0., 10., 0. 1, 1, -1000
*ELEMENT, TYPE=T2D2, ELSET=STEEL RODS *NODE PRINT
1, 1, 2 U,
3, 2, 3 RF,
5, 3, 4 *EL PRINT
*ELEMENT, TYPE=T2D2, ELSET=ALUMINUM RODS S,
2, 1, 3 ** OUTPUT FOR ABAQUS QA PURPOSES
4, 2, 4 * EL FILE
*SOLID SECTION, ELSET=STEEL RODS, MATERIAL=STEEL S,
** diameter=0.5 in. (area=0.1963 in^2) * NODE FILE
0.1963, U, RF
*MATERIAL, NAME=STEEL *END STEP
*ELASTIC
30000000, 0.3 >>Save as barras.inp
*SOLID SECTION, ELSET=ALUMINUM RODS, MATERIAL=ALUMINUM
** diameter=0.5 in. (area=0.1257 in^2)
To run the file barras.inp use the
0.1257, command:
*MATERIAL, NAME=ALUMINUM >>abaqus job=barras
*ELASTIC
11000000, 0.3
ABAQUS Field Output Report, written Sun May 24
14:19:06 2020
Element Label E11 S11
-------------------------------------------------------------
1 294.116E-06 8.82348E+03
3 -394.598E-06 -11.8379E+03
5 176.470E-06 5.29409E+03
-------------------------------------------------------------
2 -723.223E-06 -7.95545E+03
4 1.58450E-03 17.4295E+03

Example 2: Two dimensional truss-overhead hoist frame

Input file: cercha.inp *ELASTIC


200.E9, 0.3
**
*HEADING
** History data---------------------------
Two-dimensional overhead hoist frame
*STEP, PERTURBATION
SI units (kg, m, s, N)
10kN central load
1-axis horizontal, 2-axis vertical
*STATIC
*PREPRINT, ECHO=YES, MODEL=YES, HISTORY=YES
*BOUNDARY
**
101, ENCASTRE
** Model definition-----------------------
103, 2
*NODE, NSET=NALL
*CLOAD
101, 0., 0., 0.
102, 2, -10000.
102, 1., 0., 0.
*NODE PRINT
103, 2., 0., 0.
U,
104, 0.5, 0.866, 0.
RF,
105, 1.5, 0.866, 0.
*EL PRINT
*ELEMENT, TYPE=T2D2, ELSET=FRAME
S,
11, 101, 102
** OUTPUT FOR ABAQUS QA PURPOSES
12, 102, 103
*EL FILE
13, 101, 104
S,
14, 102, 104
*NODE FILE
15, 102, 105
U, RF
16, 103, 105
*END STEP
17, 104, 105
*SOLID SECTION, ELSET=FRAME, MATERIAL=STEEL
** diameter = 5mm --> area = 1.963E-5 m^2 >>Save as cercha.inp
1.963E-5, To run the file cercha.inp use the command:
*MATERIAL, NAME=STEEL >>abaqus job=cercha
Example 3: Two dimensional truss-Cantilever Beam

Input file: voladizo.inp

*HEADING 19, 4, 10
TRUSS CANTILEVER BEAM 20, 5, 11
** *SOLID SECTION, ELSET=BEAM, MATERIAL=STEEL
** This is in inches ** D=1 in, A=0.7854 in ^2
** 0.7854,
*NODE *MATERIAL, NAME=STEEL
1, 0., 0. *ELASTIC
6, 100., 0. 30.E6,
7, 0., 20. *BOUNDARY
12, 100., 20. FIXED, ENCASTRE
*NSET, NSET=FIXED *STEP, PERTURBATION
1, 7 *STATIC
*NGEN, NSET=allnodes **
1, 6 ** The load is in pounds
7, 12 **
*ELEMENT, TYPE=T2D2, ELSET=BEAM *CLOAD
1, 1, 2 6,1,200.
2, 2, 3 12,2,-100.
3, 3, 4 *EL PRINT, POSITION=AVERAGED AT NODES, SUMMARY=YES
4, 4, 5 S, E
5, 5, 6 *NODE FILE, NSET=allnodes
6, 7, 8 U, CF, RF
7, 8, 9 *OUTPUT, FIELD, VARIABLE=PRESELECT
8, 9, 10 *ELEMENT OUTPUT
9, 10, 11 S, E
10, 11, 12 *OUTPUT, HISTORY
11, 1, 8 *NODE OUTPUT, NSET=allnodes
12, 2, 9 U, CF, RF
13, 3, 10 *END STEP
14, 4, 11
15, 5, 12 >>Save as voladizo.inp
16, 6, 12 To run the file voladizo.inp use the command:
17, 2, 8
18, 3, 9
>>abaqus job=voladizo
Example 4: Two dimensional truss-three bar

Input file: truss-3bar.inp

*HEADING 2,1,3
TRUSS-3BAR 3,1,3
*PARAMETER *STEP,NAME=STEP-1,PERTURBATION
E=30E6 *STATIC
Load=-10000 *CLOAD
*NODE 4,2,<Load>
4,10.,0. *OUTPUT,FIELD,VARIABLE=PRESELECT
1,5.,10. *EL PRINT,ELSET=BARRAS
2,7.5,10. S,E,COORD
3,10.,10. *NODE PRINT
*ELEMENT,TYPE=T2D2,ELSET=BARRAS U,COORD
1,4,1 *EL FILE
2,4,2 S,E,COORD
3,4,3 *NODE FILE
*SOLID SECTION,ELSET=BARRAS,MATERIAL=MAT1 U,RF
0.1 *END STEP
*MATERIAL,NAME=MAT1
*ELASTIC >>Save as truss-3bar.inp
<E>
*BOUNDARY
To run the file truss-3bar.inp use the command:
1,1,3 >>abaqus job=truss-3bar
Example 5: Two dimensional frame-truss

Input file: frame-truss.inp

*HEADING *STEP,NAME=STEP-1,PERTURBATION
**FRAME-TRUSS *STATIC
**------------------ *CLOAD
*PREPRINT, MODEL=NO, HISTORY=NO, ECHO=NO N2,1,100
*RESTART, WRITE, FREQUENCY=5 *BOUNDARY
*NODE, NSET=LIMITES N1,ENCASTRE
1 N4,ENCASTRE
2,0.,1.,0. **-------------------------------------------------------------------
3,1.,1.,0. *NODE PRINT, NSET=TODOS,FREQ=1,SUMMARY=NO
4,1.,0.,0. COORD,U
*NSET,NSET=N1 *EL PRINT,ELSET=SPECIMEN,FREQ=1,SUMMARY=NO
1 E,S
*NSET,NSET=N2 **
2 *OUTPUT,FIELD,FREQ=1,VARIABLE=ALL
*NSET,NSET=N3 *ELEMENT OUTPUT,ELSET=SPECIMEN,POSITION=CENTROIDAL
3 E,S
*NSET,NSET=N4 *NODE OUTPUT,NSET=TODOS
4 U,RF
** **
*NSET,NSET=TODOS *OUTPUT,HISTORY,FREQ=1
N1,N2,N3,N4 *ELEMENT OUTPUT,ELSET=SPECIMEN
** E,S
*ELEMENT,TYPE=T2D2,ELSET=SPECIMEN *NODE OUTPUT,NSET=TODOS
1,1,2 U,RF
2,2,3 *NODE FILE
3,3,4 U,RF
4,1,3 *END STEP
**
*SOLID SECTION, ELSET=SPECIMEN, MATERIAL=MAT1
4.E-4
*MATERIAL, NAME=MAT1 >>Save as frame-truss.inp
*ELASTIC To run the frame-truss.inp use the command:
200E9,0.3 >>abaqus job=frame-truss
**-------------------------------------------------------------
Example 6: Frame-beam with I Section

Input file: frame-beam.inp

*HEADING **---------------------------------------------------------------
**FRAME-BEAM *STEP,NAME=STEP-1,PERTURBATION
**------------------ *STATIC
*PREPRINT, MODEL=NO, HISTORY=NO, ECHO=NO *CLOAD
*RESTART, WRITE, FREQUENCY=5 N2,1,100
*NODE, NSET=LIMITES *BOUNDARY
1 N1,ENCASTRE
2,0.,1.,0. N4,ENCASTRE
3,1.,1.,0. **---------------------------------------------------------------
4,1.,0.,0. *NODE PRINT, NSET=TODOS,FREQ=1,SUMMARY=NO
*NSET,NSET=N1 COORD,U
1 *EL PRINT,ELSET=SPECIMEN,FREQ=1,SUMMARY=NO
*NSET,NSET=N2 E,S
2 **
*NSET,NSET=N3 *OUTPUT,FIELD,FREQ=1,VARIABLE=ALL
3 *ELEMENT OUTPUT,ELSET=SPECIMEN,POSITION=CENTROIDAL
*NSET,NSET=N4 E,S
4 *NODE OUTPUT,NSET=TODOS
** U,RF
*NSET,NSET=TODOS **
N1,N2,N3,N4 *OUTPUT,HISTORY,FREQ=1
** *ELEMENT OUTPUT,ELSET=SPECIMEN
*ELEMENT,TYPE=B31,ELSET=SPECIMEN E,S
1,1,2 *NODE OUTPUT,NSET=TODOS
2,2,3 U,RF
3,3,4 *NODE FILE
4,1,3 U,RF
** *END STEP
*Beam Section, elset=SPECIMEN, material=MAT1, section=I
0.1317, 0.18, 0.091, 0.091, 0.008, 0.008, 0.0053 >>Save as frame-beam.inp
0.,0.,1. To run the frame-beam.inp use the command:
*Material, name=MAT1
*ELASTIC
>>abaqus job=frame-beam
200E9,0.3
Example 7: Cantilever-beam with rectangular section

Input file: beam.inp *STEP, PERTURBATION


*STATIC
*HEADING **
CANTILEVER BEAM ** The load is in pounds
** **
** This is in inches *CLOAD
** 6, 2,-20000.
*NODE *EL PRINT, POSITION=AVERAGED AT NODES, SUMMARY=YES
1, 0. S11, E11,
6, 100. SF,
*NGEN, NSET=allnodes *NODE FILE, NSET=allnodes
1, 6 U, CF, RF
*ELEMENT, TYPE=B31 *OUTPUT, FIELD, VARIABLE=PRESELECT
1, 1, 2 *ELEMENT OUTPUT
*ELGEN, ELSET=BEAM SF,
1, 5 *OUTPUT, HISTORY
*BEAM SECTION, SECTION=RECTANGULAR, ELSET=BEAM, *NODE OUTPUT, NSET=allnodes
MATERIAL=STEEL U, CF, RF
1., 2. *END STEP
*MATERIAL, NAME=STEEL
*ELASTIC >>Save as beam.inp
30.E6, To run the beam.inp use the command:
*BOUNDARY >>abaqus job=beam
1, ENCASTRE
5-3 THREE-DIMENSIONAL TRUSS FINITE ELEMENT FORMULATION

Let us now extend the development in Sec. 5-1 to the three-dimensional truss. Consider the global
xyz coordinate system shown in Fig. 5-10 (a). Note the position vector r which defines an arbitrary
point P. Let us write the vector r in terms of its three Cartesian components as follows

𝐫 = 𝑟𝑥 𝐢 + 𝑟𝑦 𝐣 + 𝑟𝑧 k (5-51)

where i, j, and k are the three unit vectors in the x, y, and z directions, respectively, as shown on the
figure.

As in the two-dimensional truss formulation, let us write the same vector r in terms of its
components in the rotated coordinate system x' y' z' where, with no loss of generality, the x' axis is
purposely taken to be in the same direction as r itself, as shown in Fig. 5-10 (b). Writing the vector r
in terms of its components (rx, ry, and rz) in the rotated system, we have

𝐫 = 𝑟𝑥′ 𝐢′ + 𝑟𝑦′ 𝐣′ + 𝑟𝑧′ 𝐤 ′ (5-52)

where i', j', and k' are the unit vectors in the x', y', and z' directions, respectively. Again the x' y' z'
coordinate system is commonly referred to as a local coordinate system. From Eqs. (5-51) and (5-52),
we have
𝑟𝑥 𝐢 + 𝑟𝑦 𝐣 + 𝑟𝑧 𝐤 = 𝑟𝑥′ 𝐢′ + 𝑟𝑦′ 𝐣′ + 𝑟𝑧′ 𝐤 ′ (5-53)

Forming the dot product of this last equation with i gives

Figure 5-10 Position vector r in a three-dimensional (a) global coordinate system and (b) local coordinate
system.

𝑟𝑥 = 𝑟𝑥′ 𝐢. 𝐢′ + 𝑟𝑦′ 𝐢. 𝐣′ + 𝑟𝑧′ 𝐢. 𝐤 ′ (5.54)

but
𝐢. 𝐢′ = cos(𝑥, 𝑥 ′ ) = 𝑛11 (5-55a)

𝐢. 𝐣′ = cos(𝑥, 𝑦 ′ ) = 𝑛12 (5-55b)


and

𝐢. 𝐤 ′ = cos(𝑥, 𝑧 ′ ) = 𝑛13 (5-55c)


Therefore, Eq. (5-54) becomes

𝑟𝑥 = 𝑛11 𝑟𝑥′ + 𝑛12 𝑟𝑦′ + 𝑛13 𝑟𝑧′ (5-56)

In a completely analogous manner, we also have

𝑟𝑦 = 𝑛21 𝑟𝑥′ + 𝑛22 𝑟𝑦′ + 𝑛23 𝑟𝑧′ (5-57)

and
𝑟𝑧 = 𝑛31 𝑟𝑥′ + 𝑛32 𝑟𝑦′ + 𝑛33 𝑟𝑧′ (5-58)

The last three equations may be written in matrix notation as

𝑟𝑥 𝑛11 𝑛12 𝑛13 𝑟𝑥′


[𝑟𝑦 ] = [𝑛21 𝑛22 𝑛23 ] [𝑟𝑦′ ] (5-59a)
𝑟𝑧 𝑛31 𝑛32 𝑛33 𝑟𝑧′

or more concisely as

𝐫 = 𝐓𝐫 ′ (5-59b)

Equation (5-59b) transforms the components of an arbitrary vector in the local coordinate system to
the components in the global system. Since it can be shown that the transformation matrix T is
orthogonal, we have

𝐓 𝑇 = 𝐓 −1

and
𝐫 ′ = 𝐓 −1 𝐫 = 𝐓 𝑇 𝐫 (5-60)
where

𝑛11 𝑛21 𝑛31


𝐓 = [𝑛12
𝑇 𝑛22 𝑛32 ]
𝑛13 𝑛23 𝑛33

We are now in a position to develop the global element stiffness matrix from the local one for the
three-dimensional truss. First of all, we must extend the notation to include the z and z' components of
the nodal displacements and nodal forces. Let us represent the three components of displacement as u,
v, w and the three components of nodal forces as U, V, W. The subscripts i and j are used as usual
to denote the two respective nodes on each element. A prime (') denotes that the variable in question
is referenced to the local system (i .e., the x' y' z' coordinate system), while the absence of a prime
refers to the global system (i.e., the xyz coordinate system). Figure 3-12 may help to clarify the
notation; only nodal displacements are shown, since the figure showing nodal forces would look
identical except that uppercase letters would be used to denote the forces.

As in the development for the two-dimensional truss, the two nodal displacements in the x' direction,
u'i and u'j are related to the two nodal forces in the same direction, U'i and U'j, by
𝐴𝐸
𝑈𝑖′ = (𝑢𝑖′ − 𝑢𝑗′ ) (5-61a)
𝐿
and
𝐴𝐸
𝑈𝑗′ = (𝑢𝑗′ − 𝑢𝑖′ ) (5-61b)
𝐿

where A represents the cross-sectional area of the element, E the modulus of elasticity, and L the
element length. Now L may be computed from

𝐿 = √(𝑥𝑗 − 𝑥𝑖 )2 + (𝑦𝑗 − 𝑦𝑖 )2 + (𝑧𝑗 − 𝑧𝑖 )2 (5-62)

For the truss element only axial loads are allowed, and so we have

𝑉𝑖′ = 𝑊𝑖′ = 0 (5-61c)


and

𝑉𝑗′ = 𝑊𝑗′ = 0 (5-61d)

Figure 5-11 Nodal displacements in xyz and x'y'z' coordinate systems.

In other words, all joints are assumed to be pinned. Let us write Eqs. (5-61) in matrix form as

1 0 0 1 0 0  ui'  U i' 
0    
 0 0 0 0 0  vi'  Vi ' 
AE  0 0 0 0 0 0  wi'  Wi ' 
      (5-63a)
L  1 0 0 1 0 0  u 'j  U 'j 
0 0 0 0 0 0  v 'j  V j' 
    
 0 0 0 0 0 0 w'j  W j' 

or much more concisely as


′ ′ ′
𝐊 𝑒 𝐚𝑒 = 𝐟 𝑒 (5-63b)

where Ke' represents the local element stiffness matrix, ae' the local element nodal displacement vector,
and fe' the local element nodal force vector.
Let us now determine the global element stiffness matrix. From the development in Sec. 5-1, it
should be obvious that

𝐚𝑒 = 𝐑𝐚𝑒
and

𝐟 𝑒 = 𝐑𝐟 𝑒
where

n11 n21 n31 0 0 0


n 0 
 12 n22 n32 0 0
TT 0  n13 n23 n33 0 0 0
R T
  (5-64)
0 T  0 0 0 n11 n21 n31 
0 0 0 n12 n22 n32 
 
 0 0 0 n13 n23 n33 

As usual, ae and fe, respectively, denote the vector of nodal displacements and the vector of nodal
forces for element e, referred to the global system. These are given exp1icitly by

a e  ui vi wi uj vj wj 
T
(5-65)

and


f e  U i Vi Wi U j V j W j 
T
(5-66)

Equation (5-63b) becomes



𝐊 𝑒 𝐑𝐚𝑒 = 𝐑𝐟 𝑒 (5-67)

which may be premultiplied by R-1 (which equals RT) to give

𝐊 𝑒 𝐚𝑒 = 𝐟 𝑒 (5-68)

where Ke represents the global element stiffness matrix and is obviously defined by

𝐊 𝑒 = 𝐑𝑇 𝐊 𝑒 𝐑 (5-69)

lf Ke is determined explicitly, one gets

 n112 n21n11 n31n11  n112  n21n11  n31n11 


 
 n11n21
2
n21 n31n21  n11n21  n212
 n31n21 
AE  n11n31 n21n31 2
n31  n11n31  n21n31  n312 
K 
e
  (5-70)
L   n112  n21n11  n31n11 n112 n21n11 n31n11 
 n n  n212
 n31n21 n11n21 2
n21 n31n21 
 11 21

 n11n31  n21n31  n31 
2 2
n11n31 n21n31 n31
Interestingly, the global element stiffness matrix Ke is a function of only three easily calculated
direction cosines, namely, n11, n21, and n31, in addition to A, E, and L. By definition of the direction
cosines, we have
𝑛11 = cos(𝑥, 𝑥 ′ ) (5-71a)

𝑛21 = cos(𝑦, 𝑥 ′ ) (5-71b)

𝑛31 = cos(𝑧, 𝑥 ′ ) (5-71c)

Let us define the vector rij which runs from node i to j in the global system as

𝐫𝑖𝑗 = (𝑥𝑗 − 𝑥𝑖 )𝐢 + (𝑦𝑗 − 𝑦𝑖 )𝐣 + (𝑧𝑗 − 𝑧𝑖 )𝐤 (5-72)


But ∣rij∣ = L, and
𝐫 𝐫𝑖𝑗
𝐢′ = ∣𝐫𝑖𝑗∣ = (5-73)
𝑖𝑗 𝐿
so we have
𝑥𝑗−𝑥𝑖 𝑦𝑗 −𝑦𝑖 𝑧𝑗−𝑧𝑖
𝐢′ = ( )𝐢 + ( )𝐣 +( )𝐤 (5-74)
𝐿 𝐿 𝐿

But dotting both sides of Eq. (5-74) with i gives

𝑥𝑗 − 𝑥𝑖
𝐢. 𝐢′ = = 𝑛11
𝐿

where i·i' = cos(x,x')=n11. Similarly dotting Eq. (5-74) with j and k gives

𝑦𝑗 − 𝑦𝑖
𝐣. 𝐢′ = = 𝑛21
𝐿
and
𝑧𝑗 − 𝑧𝑖
𝐤. 𝐢′ = = 𝑛31
𝐿

In summary, the three direction cosines n11, n21, and n31 are given by
𝑥𝑗 −𝑥𝑖
𝑛11 = cos(𝑥, 𝑥 ′ ) = (5-75a)
𝐿

𝑦𝑗 −𝑦𝑖
𝑛21 = cos(𝑦, 𝑥 ′ ) = (5-75b)
𝐿

𝑧𝑗 −𝑧𝑖
𝑛31 = cos(𝑧, 𝑥 ′ ) = (5-75c)
𝐿

where L, of course, is the length of the element given by Eq. (5-62). For any element, the global nodal
coordinates are known since they are an integral part of the input to any finite element program.
Therefore, the global element stiffness matrix Ke is easily computed from Eq. (5-70).

The assemblage process remains the same as described in Sec. 5-1 except that each of the four 2 x
2 submatrices in Ke is now a 3 x 3 submatrix. Prescribed displacement boundary conditions are applied
in the usual manner. External nodal forces are also applied in the same manner as described in Sec. 5-
1 for the two dimensional truss. The resulting assemblage system equation

Ka = f (5-76)
is solved for the nodal displacements, which in tum can be used to solve for the internal axial forces
and stresses in each element. The procedure is summarized below. The elongation δ of an element is
given by
δ = 𝑢𝑗′ − 𝑢𝑖′ (5-77)
where from Eq. (5-60), we have
𝑢𝑖′ = 𝑛11 𝑢𝑖 + 𝑛21 𝑣𝑖 + 𝑛31 𝑤𝑖 (5-78a)

and
𝑢𝑗′ = 𝑛11 𝑢𝑗 + 𝑛21 𝑣𝑗 + 𝑛31 𝑤𝑗 (5-78b)

Since the direction cosines are readily calculated with the help of Eq. (5-75), and since the nodal
displacements ui, vi, wi, uj, vj, and wj are now known (i.e., from the solution for a in Ka = f), the
elongation δ is easily found for each element. The axial strains ɛ, stresses σ, and forces F within each
element can then be determined by using Eqs. (5-43) to (5-45). Negative values of ɛ, σ, and F denote
compression (when δ < 0). This completes the formulation of the three-dimensional truss problem.

Example 1: Three dimensional truss

Now that you have performed analysis on 2D trusses, it is time to move on to a 3D truss structure.
Listed below are tables showing the required boundary conditions and loads at the specified nodes.
For the material properties use steel (E=30. x106).

Node # Fixed Displacements


1 u,v,w
7 u,v,w
13 u,v,w
18 u,v,w
Node # Direction of Load Load
11 w -100
12 w -200
22 w -100
12 u 75
6 u -75

Input file: 3Dtruss.inp


*HEADING 19,-12.0,25.0,20.0
3D TRUSS PROBLEM, Written by: E. A. Bonifaz 20,-9.0,50.0,20.0
*PREPRINT, MODEL=NO, HISTORY=NO, ECHO=NO 21,-6.0,75.0,20.0
** First introduce the nodal positions 22,-3.0,100.0,20.0
*NODE, NSET=NODES1 ** Then introduce the connectivity table
1,12.0,0.0,0.0 *ELEMENT, TYPE=T3D2, ELSET=TRUSS3D
2,12.0,25.0,0.0 1,1,2
3,9.0,50.0,2.5 2,2,3
4,6.0,75.0,5.0 3,3,4
5,3.0,100.0,7.5 4,4,5
6,0.0,125.0,10.0 5,5,6
7,12.0,0.0,20.0 6,1,7
8,12.0,25.0,20.0 7,1,8
9,9.0,50.0,20.0 8,2,8
10,6.0,75.0,20.0 9,2,9
11,3.0,100.0,20.0 10,3,9
12,0.0,125.0,20.0 11,3,10
13,-12.0,0.0,0.0 12,4,10
14,-12.0,25.0,0.0 13,4,11
15,-9.0,50.0,2.5 14,5,11
16,-6.0,75.0,5.0 15,5,12
17,-3.0,100.0,7.5 16,6,12
18,-12.0,0.0,20.0 17,7,8
18,8,9 59,11,22
19,9,10 60,1,18
20,10,11 61,2,19
21,11,12 62,3,20
22,13,14 63,4,21
23,14,15 64,5,22
24,15,16 **Next we enter the material description.
25,16,17 *SOLID SECTION, ELSET=TRUSS3D, MATERIAL=STEEL
26,17,6 0.1
27,13,18 *MATERIAL, NAME=STEEL
28,13,19 *ELASTIC
29,14,19 30.0E6
30,14,20 **Next we impose the boundary conditions
31,15,20 *BOUNDARY
32,15,21 1,1,3
33,16,21 13,1,3
34,16,22 7,1,3
35,17,22 18,1,3
36,17,12 **Type of analysis and load description
37,18,19 *STEP, PERTURBATION
38,19,20 *STATIC
39,20,21 *CLOAD
40,21,22 11,3,-100
41,22,12 12,3,-200
42,1,13 22,3,-100
43,1,14 12,1,75
44,2,14 6,1,-75
45,2,15 **Finally, specify the content of the output files
46,3,15 *EL PRINT,SUMMARY=YES
47,3,16 S11,E11
48,4,16 *EL FILE,ELSET=TRUSS3D
49,4,17 S,E
50,5,17 *NODE FILE,NSET=NODES1
51,7,18 U,CF,RF
52,7,19 *RESTART, WRITE
53,8,19 *END STEP
54,8,20
55,9,20 >>Save as 3Dtruss.inp
56,9,21 To run the 3Dtruss.inp use the command:
57,10,21 >>abaqus job=3Dtruss
58,10,22

REFERENCES
1. Stassa, F. L., Applied Finite Element Analysis for Engineers. HRW Series in Mechanical
Engineering. CBS Publishing -Japan Ltd. 1985.
6
WEIGHTED RESIDUAL FORMULATIONS

6-1 INTRODUCTION

In Chapter 5 the direct approach was introduced when the two-dimensional truss problem was
formulated. It was pointed out that alternate, more indirect approaches exist that may be used to
develop finite element models [1]. Some of these methods of formulation are introduced in this chapter.
The reader will soon come to appreciate that virtually all problems that are describable by ordinary
and partial differential equations can be solved by the finite element method. Moreover, linear and
nonlinear problems may be solved in nearly the same way, although nonlinear problems generally
require an iterative solution. In this chapter only steady-state problems are studied. By steady-state we
mean that the field variable is a function of spacial coordinates only and not a function of time.
Problems that allow for time-varying field variables are said to be unsteady, transient, dynamic, or
time-dependent. In some texts transient problems are referred to as propagation problems [1]. In
structural and stress analysis, time-independent problems are referred to as static or equilibrium
problems, whereas time-dependent problems are almost exclusively referred to as dynamic.

In all steady-state, static, or equilibrium problems, a system of algebraic equations results that is
always in the form
Ka = f

It should be recalled that the truss model resulted in a system of equations in exactly this form.
Furthermore, the vector or column matrix a always contains the nodal unknowns, which are really the
values of the field variables at the nodal points. It cannot be emphasized enough that in structural
models (without bending), these unknowns are the nodal displacements; in thermal models, they are
the nodal temperatures; and in fluid flow problems, they are the nodal velocities and pressures. The
main purpose of a finite element analysis is to determine the values of the field variable(s) at the node
points. Other quantities such as the stresses or heat-flows may then be determined in subsequent
calculations. This chapter is further restricted to the study of only one-dimensional problems. This
means that the field variable of interest is a function of only one variable, e.g., x.

Oftentimes exact solutions to differential equations are not easily obtained. In fact, quite frequently,
it is not even possible to obtain an exact solution with currently available mathematical techniques.
This is particularly true when the geometry is irregular, or the properties vary spatially, or perhaps, the
properties are a function of the field variable. Problems of the latter variety are said to be nonlinear.
Engineers, mathematicians, and applied scientists find approximate solution methods indispensable in
these cases. Among the approximate solution techniques that are useful to a thorough understanding
of how the finite element method works are the following: The Ritz method, the variational, or
Rayleigh-Ritz, method, and the weighted-residual method. Each of these is said to be an integral
method because we work with an integral form of the problem statement instead of the governing
differential equation directly.
6-2 THE METHOD OF WEIGHTED RESIDUALS

General concepts

The method of weighted residuals provides an analyst with a very powerful approximate solution
procedure that is applicable to a wide variety of problems. It is the existence of the various weighted-
residual methods that makes it unnecessary to search for variational formulations to nonstructural
problems in order to apply the finite element method to these problems. The most popular weighted-
residual methods are referred to as (1) point collocation, (2) subdomain collocation, (3) least squares,
and (4) Galerkin. Of these four methods, the Galerkin method has the widest use in finite element
analysis for reasons that will become apparent later [1].

Before actually illustrating the weighted-residual method with a specific problem, let us present a
general framework around which the various methods are developed. So as not to obscure the basic
ideas behind the approximate solution methods, let us restrict the present discussion to a single
governing differential equation with only one independent variable. Let us represent the governing
equation as

𝑓[𝑇(𝑥)] = 0 in Ω (6-1)

where T represents the function sought (e. g., temperature) that is a function of x only. The symbol Ω
represents the domain of the region governed by Eq. (6-1). In addition, let us specify the boundary
conditions symbolically in the form

g1 [𝑇(𝑥)] = 0 on Γ1 (6-2a)
and

g 2 [𝑇(𝑥)] = 0 on Γ2 (6-2b)

etc., where Γ1 and Γ2 include only those parts of the domain Ω that are on the boundary. Figure 6-1
may help to clarify the notation.

Figure 6-1 Schematic of one-dimensional problem domain Ω with two global boundaries Γ 1 and Γ2.

Let us approximate the solution to Eqs. (6-l) and (6-2) with the approximate function T' where

n
T '  T ' ( x; a1 , a2 ,..., an )   ai Ni ( x) (6-3)
i 1

which has one or more unknown (but constant) parameters a1 , a2 , ….an and that satisfies the boundary
conditions given by Eqs. (4-2) exactly. Note that the prime (') denotes an approximate solution (not a
derivative). The functions Ni(x) are referred to as trial functions. The problem reduces to having to
make somewhat judicious choices for these trial functions and solving for the parameters a1, a2, ... an.
In general, if a sequence of approximations could be made, such as
𝑇 ′ = 𝑇 ′ (𝑥; 𝑎1 ) = 𝑎1 𝑁1 (𝑥) (6-4a)

𝑇 ′ = 𝑇 ′ (𝑥; 𝑎1 , 𝑎2 ) = 𝑎1 𝑁1 (𝑥) + 𝑎2 𝑁2 (𝑥) (6-4b)

𝑇 ′ = 𝑇 ′ (𝑥; 𝑎1 , 𝑎2 , 𝑎3 ) = 𝑎1 𝑁1 (𝑥) + 𝑎2 𝑁2 (𝑥) + 𝑎3 𝑁3 (𝑥) (6-4c)

then presumably better accuracy could be obtained with each successive higher-order approximation.
Equations (6-4) are referred to as the first-order, second-order, and third-order approximations,
respectively. The reader is referred to Becker, Carey, and Oden [2] for a discussion on the necessary
properties of the trial functions. Suffice it to say here that they must be continuous and differentiable
up to the highest order present in the integral form of the governing equation.

It should not be surprising that if T' given in Eq. (6-3) is substituted for T in Eq. (6-1), the governing
equation will not be satisfied exactly. Instead of getting f(T') equal to zero, we get a residual R(x; a1,
a2,...,an). Mathematically, we may write

𝑓[𝑇 ′ (𝑥; 𝑎1 , 𝑎2 , … , 𝑎𝑛 )] = 𝑅(𝑥; 𝑎1 , 𝑎2 , … , 𝑎𝑛 ) (6-5)

where the notation is supposed to indicate that the residual R is a function of x and the parameters a1,
a2, etc. The exact solution results when the residual R is zero for all points in the domain Ω. For the
approximate solution methods, the residual is not in general zero everywhere in Ω although it may be
zero at some selected points.

The method of weighted residuals requires that the parameters a1, a2, a3, .... an be determined by
satisfying
 wi ( x)R( x; a1, a2 ,..., an )dx  0 for i  1, 2,..., n

(6-6)

where the functions wi(x) are the n arbitrary weighting functions. The choice of the weighting functions
is left largely up to personal preference, but four particular functions are used most often.

Galerkin

In the Galerkin weighted-residual method, the trial functions Ni(x) themselves are used as the
weighting functions, or
𝑤𝑖 (𝑥 ) = 𝑁𝑖 (𝑥) (6-7)

For the Galerkin method Eq. (6-6) becomes


Ni ( x) R( x; a1 , a2 ,..., an )dx  0 for i  1, 2,..., n (6-8)

Because there is one trial function for each unknown parameter, Eq. (6-8) rea1ly gives n such equations
that, when solved, yield the values of the unknown parameters, a1, a2, a3, .... an. It should be obvious
that again the values obtained for the ai’s are dependent on the choice of trial functions. The Galerkin
method is seen to be quite simple.

Example 6-1

Solve the ordinary linear differential equation


𝑑2 𝑇
+ 1000𝑥 2 = 0 0≤𝑥≤1 (6-9)
𝑑𝑥 2

subject to the boundary conditions


T(0) = 0 and T(1) = 0 (6-10)

by using the Galerkin method with the trial function

𝑁1 (𝑥 ) = 𝑥(1 − 𝑥 2 ) (6-11)
Solution

The residual R by definition may be computed from

𝑑2𝑇 ′
𝑅= + 1000𝑥 2
𝑑𝑥 2
and since

𝑑2𝑇 ′
= −6𝑎1 𝑥
𝑑𝑥 2

we have
𝑅(𝑥; 𝑎1 ) = −6𝑎1 𝑥 + 1000𝑥 2 (6-12)

Substituting R from Eq. (6-12) and N1(x) from Eq. (6-11) into Eq. (6-8) for i = 1 only gives
1
∫ 𝑥 (1 − 𝑥 2 )(−6𝑎1 𝑥 + 1000𝑥 2 )𝑑𝑥 = 0
0

Simplifying the integrand, integrating, evaluating, and solving for a1 yields

5000
𝑎1 =
48

Therefore, an approximate solution by the Galerkin method is given by

5000
𝑇𝐺′ = 𝑥(1 − 𝑥 2 ) (6-13)
48

Comparison with the Exact Solution

If the problem posed in the Example 6-1 is solved exactly, we get


1000
𝑇 (𝑥 ) = 𝑥(1 − 𝑥 3 ) (6-14)
12

Note the exponent of 3 on the x in the parentheses; our approximate solution was assumed to be of the
form
𝑇 ′ (𝑥 ) = 𝑎1 𝑥(1 − 𝑥 2 ) (6-15)
which is slightly different than the exact solution above. The exact solution from Eq. (6-14) is
compared to the Galerkin solution in Fig. 6-2. It is seen a difference between the exact and the
approximate Galerkin solution. However, in Sec. 6-3 the Galerkin method is adapted to a finite element
formulation for a solution to the same problem. The fundamental difference between what we have
done so far and what we will be doing next lies in the type of trial functions that are assumed. Recall
that trial functions were chosen that applied globally to the entire domain being analyzed. From this
point onward, trial functions will be chosen that are to be applied locally (over each and every finite
element). These trial functions will be said to be piecewise continuous. Piecewise continuous trial
functions in this text will be referred to as shape functions.

45
40
35
Temperature, T

30
25
20 Exact Solution

15 Galerkin

10
5
0
0 0.2 0.4 0.6 0.8 1
Distance along rod, x
Figure 6-2 Comparison of approximate Galerkin solution to problem posed in Example 6-1 with the exact
solution.

This important step is fundamental to the finite element method and has associated with it two
important implications: (1) the piecewise continuous trial functions or shape functions need not satisfy
the physics of the problem since they are to be applied locally in every element, not globally, and (2)
the global boundary conditions need not be satisfied. This is not to say that the shape functions have
no restrictions placed on them.

6-3 THE GALERKIN FINITE ELEMENT METHOD

In Sec. 6-2, the Galerkin weighted-residual method was introduced. Global trial functions were
used; these functions applied to the entire domain. In this section we wish to repeat the solution in the
context of the finite element method. The implication is that piecewise continuous trial functions or
shape functions (interpolation polynomials) will be employed for each element.

Consider the domain that includes the interval a ≤x ≤ b as shown in Fig. 6-3. Let us divide this
interval into M subintervals as shown. Actually, this process is called discretization and a typical
subinterval from xj to xk is really a finite element and is simply denoted as element e. The ends of the
subinterval are the locations of the two nodes, nodes j and k.
Figure 6-3 Typical one-dimensional problem domain discretized into M elements with N nodes with an assumed
linear variation of the field variable y(x) over a typical element e with nodes j and k.

Since the element e has two nodes, a linear interpolating polynomial may be used to describe the
behavior of the field variable, e.g., y(x), over the element, or

𝑦 𝑒 = 𝑚𝑥 + 𝑏 (6-16)

The superscript (e) on the field variable serves to remind us that the approximation applies only over
element e and not globally. But at x = xj we must have ye = yj, and at x = xk , we must have ye = yk.
Therefore, applying these two conditions to Ec. (6-16) yields

𝑦𝑗 = 𝑚𝑥𝑗 + 𝑏 (6-17a)

𝑦𝑘 = 𝑚𝑥𝑘 + 𝑏 (6-17b)

that may be solved for m and b to give


𝑦𝑘 −𝑦𝑗
𝑚= (6-18a)
𝑥𝑘 −𝑥𝑗
and
𝑦 −𝑦
𝑏 = 𝑦𝑗 − ( 𝑥𝑘 −𝑥𝑗 ) 𝑥𝑗 (6-18b)
𝑘 𝑗

Substituting these expressions for m and b in Eq. (6-16) and some rearrangement gives the following:

𝑥 −𝑥 𝑥−𝑥𝑗
𝑦 𝑒 = (𝑥 𝑘−𝑥 ) 𝑦𝑗 + (𝑥 ) 𝑦𝑘 (6-19)
𝑘 𝑗 𝑘 −𝑥𝑗

It is observed that Eq. (6-19) is of the form

𝑦 𝑒 = 𝑎1 𝑁1 (𝑥 ) + 𝑎2 𝑁2 (𝑥) (6-20a)
or
𝑦 𝑒 = 𝑎𝑗 𝑁𝑗 (𝑥) + 𝑎𝑘 𝑁𝑘 (𝑥) (6-20b)
or
𝑦 𝑒 (𝑥) = 𝑦𝑗 𝑁𝑗 (𝑥 ) + 𝑦𝑘 𝑁𝑘 (𝑥) (6-20c)
where
𝑎1 = 𝑎𝑗 = 𝑦𝑗 (6-21a)

𝑎2 = 𝑎𝑘 = 𝑦𝑘 6-21b)
𝑥 −𝑥
𝑁1 (𝑥 ) = 𝑁𝑗 (𝑥 ) = 𝑥 𝑘−𝑥 (6-22a)
𝑘 𝑗
and
𝑥−𝑥𝑗
𝑁2 (𝑥 ) = 𝑁𝑘 (𝑥 ) = 𝑥 (6-22b)
𝑘 −𝑥𝑗

The functions Nj(x) and Nk(x) are referred to as the shape functions for the element. We see that
instead of using higher-order polynomials that apply globally to the entire domain, we will now use
two first-order shape functions that apply only to a small subinterval or element. These particular shape
functions are shown in Fig. 6-4.

Figure 6-4 Typical element e showing the two shape functions and the resulting variation of the field variable
ye(x) (linear between nodes j and k).

Note that they are both linear and at x = xj, we have Nj = 1 and Nk = 0, and at x = xk we have Nj = 0
and Nk = 1. Equation (6-20b) may be written in matrix form as follows:

𝑦 𝑒 = 𝐍𝐚𝑒 (6-23)
where by definition
𝑁 = [𝑁𝑗 (𝑥 ) 𝑁𝑘 (𝑥)] (6-24a)
and
𝑎𝑗
𝑎𝑒 = [𝑎 ] (6-24b)
𝑘

It should be emphasized that the vector ae contains the values of the field variable at the two nodes.
The idea is to use Eq. (6-23) to represent y(x) over each element in the discretized domain and to find
the values of the ai’s. The set of all ai's for the entire domain represents the finite element solution to
the original problem.

Because the function y(x) itself given by Eq. (6-16) is continuous and the first derivatives are not
necessarily continuous (from element to element), the function y(x) is said to be C0-continuous. A
function with continuous first derivatives is said to be C1-continuous, etc. In a similar fashion, the two
shape functions given by Eq. (6-22) are also said to be C0-continuous or to possess C0-continuity.
These rather abstract ideas are best illustrated by tackling a numerical example, which again, for
comparison purposes, is taken to be the problem posed in Example 6-1.

From Eqs. (6-20) to (6-22), we see that at x = xj we have ye(xj) = yj, and at x = xk we have ye(xk) =
yk, and two piecewise continuous linear trial functions are to be used. The functions Nj(x) and Nk(x)
are referred to as shape functions.

Example 6-2

Resolve the problem posed in Example 6-1 (approximately) by the Galerkin finite element method by
using six equally spaced nodes.

Solution

Recall that the Galerkin method requires that the trial functions themselves be used as the weighting
functions. The integral of the weighted residual R in this case is written as


Ni ( x) R( x; a1 , a2 ,..., an )dx  0 (6-25)

But instead of using trial functions Ni(x) that are applied globally to the entire x domain and that satisfy
the boundary conditions automatically (and exactly), let us try to represent the solution Te over a small
interval (really a finite element), say from xj to xk, as

𝑇 𝑒 = 𝑁𝑗 (𝑥 )𝑇𝑗 + 𝑁𝑘 (𝑥)𝑇𝑘 (6-26)

where Tj and Tk are analogous to the unknown parameters, a1 and a2, in the general, weighted-residual
method. Moreover, if Nj(x) and Nk(x), in addition to being linear, are such that when x = xj we have Nj
= l and Nk = 0, and when x = xk we have Nj = 0 and Nk = 1, then in fact we have appropriate shape
functions. The reader should verify that for Nj(x) and Nk(x) given in Eq. (6-22) this is in fact the case.
In summary, in one-dimensional C0-continuous problems with two nodes per element, the shape
functions must be linear because there are two nodal points per element and they must satisfy the
following conditions:
𝑁𝑗 (𝑥𝑗 ) = 𝑁𝑘 (𝑥𝑘 ) = 1 (6-27a)
and
𝑁𝑗 (𝑥𝑘 ) = 𝑁𝑘 (𝑥𝑗 ) = 0 (6-27b)

In the Galerkin method, the two shape functions Nj and Nk are the weighting functions. From Eq.
(6-8), we may write the following weighted-residual equations:

 
M


xk
N j ( x) Re ( x; T j , Tk ) dx  0 (6-28a)
xj
e1
and
 
M


xk
N k ( x) Re ( x; T j , Tk ) dx  0 (6-28b)
xj
e1

where e denotes the element and M is the number of elements (or intervals into which the domain is
divided). A typical finite element goes from xj (where the temperature is Tj) to xk (where the
temperature is Tk). If matrix notation is used and NT is defined as
𝑁𝑗 (𝑥)
𝐍𝑇 = [ ]
𝑁𝑘 (𝑥)

then Eqs. (6-28) may be written in one equation as

 
M


xk
NT Re ( x; T j , Tk ) dx  0 (6-29)
xj
e1

But the residual Re for a typical element e in the problem at hand is, by definition,

𝑑2 𝑇 𝑒
𝑅𝑒 = + 1000𝑥 2 (6-30)
𝑑𝑥 2

where the approximate solution T, applicable only to element e, has been substituted for the exact
temperature T in the governing equation and the result set equal to the residual Re for the element.
Therefore, Eq. (6-29) becomes
Td T 2
M 2 e

 
xk
N 
 dx 2  1000x dx  0 (6-31)
e1
xj
 

Before breaking this integral into two separate integrals, let us agree to drop the summation sign. The
reason for this is simple: it represents the assemblage step that it is hoped is becoming routine. In other
words, it is preferred to work with a typical element (element e) and at some later stage we will put
the pieces back together. Breaking Eq. (6-31) into two separate integrals gives

d 2T e
 dx   NT (1000x 2 ) dx  0
xk xk
NT 2
(6-32)
xj dx x j

If the first integral is integrated by parts we get

xk
dT e dNT dT e
 dx   NT (1000x 2 )dx  0
xk xk
T
N (6-33)
dx xj
xj dx dx xj

But we have already agreed to write Te in terms of two (linear) shape functions, Nj and Nk, and the two
nodal temperatures, Tj and Tk, as
𝑇 𝑒 = 𝐍𝐚𝑒 (6-34)
where
𝐍 = [𝑁𝑗 𝑁𝑘 ] (6-35)
and
𝑇𝑗
𝐚𝑒 = [ ] (6-36)
𝑇𝑘

Using Eq. (6-34) in the integral terms only in Eq. (6-33) gives

xk
dT e dNT dN e
 a dx   NT (1000x 2 )dx  0
xk xk
T
N (6-37)
dx xj
x j dx dx xj
The reason for not using Eq. (6-34) in the integrated term is that by not doing so, we have allowed a
mechanism for imposing gradient boundary conditions. But ae is not a function of x and may be pulled
out of the integral from the right to give a rather simple equation of the form

𝐊 𝑒 𝐚𝑒 = 𝐟 𝑒 (6-38)
where
dNT dN
Ke  
xk
dx (6-39)
xj dx dx
and
xk
dT e
  NT (1000x 2 )dx  0
xk
f N
e T
(6-40)
dx xj
xj

Equation (6-39) for the element stiffness matrix and Eq. (6-40) for the element nodal force vector are
the so-called element characteristics. Note that it is not necessary to distinguish between the local and
global element stiffness matrices since a local coordinate system was not used in the development of
these equations. Moreover, the function T(x) is a scalar function (such as temperature) and hence no
transformation from a local to a global coordinate system would be necessary before the assemblage
step.

It should be noted that Ke is a symmetric 2 x 2 matrix, whereas fe is a 2 x 1 matrix (or column


vector). Obviously, ae is also a 2 x 1 matrix. It is very easy to evaluate Ke given by Eq. (6-39) if the
expressions for Nj(x) and Nk(x) from Eq. (6-22) are used. The final result may be verified to be

1 1 −1
𝐊𝑒 = 𝑥 [ ] (6-41)
𝑘 −𝑥𝑗 −1 1

The integrated term in Eq. (6-40) is evaluated as follows:

𝑑𝑇 𝑥𝑘 𝑑𝑇(𝑥𝑘 ) 𝑑𝑇(𝑥𝑗 )
𝐍𝑇 | = 𝐍 𝑇 (𝑥𝑘 ) − 𝐍 𝑇 (𝑥𝑗 ) (6-42)
𝑑𝑥 𝑥𝑗 𝑑𝑥 𝑑𝑥

But from Eqs. (6-22) and (6-24a) we have

0
𝐍 𝑇 (𝑥𝑘 ) = [ ]
1
and
1
𝐍 𝑇 (𝑥𝑗 ) = [ ]
0

Therefore, the integrated term becomes


𝑥𝑘
𝑑𝑇 𝑑𝑇(𝑥𝑘 ) 0 𝑑𝑇(𝑥𝑗 ) −1
𝐍𝑇 | = [ ]− [ ] (6-43)
𝑑𝑥 𝑥𝑗 𝑑𝑥 1 𝑑𝑥 0

The evaluation of the remaining part of fe, namely the integral contribution, is very tedious but
routine. The reader should show that the final result is given by

𝑥𝑘 1000/12 𝑥𝑘4 − 4𝑥𝑘 𝑥𝑗3 + 3𝑥𝑗4


∫𝑥 1000𝑥 2 𝑁 𝑇 𝑑𝑥 = [ 4 ] (6-44)
𝑗 𝑥𝑘 −𝑥𝑗 3𝑥𝑘 − 4𝑥𝑗 𝑥𝑘3 + 𝑥𝑗4
Figure 6-5 Discretized problem domain for problem posed in example 6-2.

Before making use of these results in carrying out the numerical computations, we must discretize the
global domain into five equally sized elements that require six equally spaced nodes, as shown in Fig.
6-5. It proves to be very convenient to organize the nodal coordinate and element connectivity data as
shown in Tables 6-1 and 6-2, respectively. Note that a material set flag is not used because only one
material is present in the model.

By using the data in these tables it is easy to compute the element stiffness matrices K(1), K(2), etc.,
and the element nodal force vectors f(1), f(2), etc. For example, for K(1) we have

1 1 −1 1 1 −1
𝐊 (1) = [ ]= [ ]=[ 5 −5]
𝑥2 − 𝑥1 −1 1 0.2 −1 1 −5 5

Table 6-1 Nodal Coordinate Data for Example 6-2

Node number x

1 0.0
2 0.2
3 0.4
4 0.6
5 0.8
6 1.0

Table 6-2 Element Data for Example 6-2

Element Nodes connected


number i j

1 1 2
2 2 3
3 3 4
4 4 5
5 5 6

Obviously, we also have

5 −5]
𝐊 (1) = 𝐊 (2) = 𝐊 (3) = 𝐊 (4) = 𝐊 (5) = [ (6-45a)
−5 5
for f(1), we have

0 𝑑𝑇(𝑥2 ) −1 𝑑𝑇(𝑥1 ) 1000 (0.2)2 − 4(0.2)(0.0)3 + 3(0.0)4


𝐟 (1) = [ ] +[ ] + [ ]
1 𝑑𝑥 0 𝑑𝑥 (12)(0.2) 3(0.2)4 − 4(0.0)(0.2)3 + (0.0)4
or

0 𝑑𝑇(𝑥2 ) −1 𝑑𝑇(𝑥1 ) 0.667


𝐟 (1) = [ ] +[ ] +[ ]
1 𝑑𝑥 0 𝑑𝑥 2.0
Similarly, we find

0 𝑑𝑇(𝑥3 ) −1 𝑑𝑇(𝑥2 ) 7.3


𝐟 (2) = [ ] +[ ] +[ ]
1 𝑑𝑥 0 𝑑𝑥 11.3

0 𝑑𝑇(𝑥 ) −1 𝑑𝑇(𝑥 ) 22.0


𝐟 (3) = [ ] 𝑑𝑥 4 + [ ] 𝑑𝑥 3 + [ ] (6-45b)
1 0 28.6
0 𝑑𝑇(𝑥5 ) −1 𝑑𝑇(𝑥4 ) 44.7
𝐟 (4) = [ ] +[ ] +[ ]
1 𝑑𝑥 0 𝑑𝑥 54.0

0 𝑑𝑇(𝑥6 ) −1 𝑑𝑇(𝑥5 ) 75.4]


𝐟 (5) = [ ] +[ ] +[
1 𝑑𝑥 0 𝑑𝑥 87.3

The assemblage of the element stiffness matrices to form Ka follows the same line of reasoning
presented in Chapter 5 where the truss was studied. However, the situation here is even simpler because
the submatrices in the Ke's are of size 1 x 1, or simply scalars. The reader is reminded that the primary
factor in deciding where the element stiffness matrix Ke for element e is added into the assemblage
stiffness matrix Ka is the two global node numbers (I and J) associated with the element. This
assemblage step was summarized in Chapter 5 for elements containing only two nodes (such as the
present problem) in Eq. (5-31). The details of the assemblage of the stiffness matrices are omitted
below, but the reader should show the missing steps if there is still some doubt as to how this important
step is accomplished.

The assemblage of the nodal force vectors to form fa is done in a completely analogous manner,
except that it is even simpler because only a column matrix is involved. If the element nodal force
vector is imagined to be partitioned as
𝑓𝑒
𝐟 𝑒 = [ 𝐼𝑒 ] (6-46a)
𝑓𝐽

then the assemblage of these to form fa may be summarized by the following equation:

M
fna  fne (6-46b)
e 1

where M is the maximum number of elements used in the model and 𝐟𝑛𝑒 is taken to be zero if element
e does not contain node n. With this algorithm, the reader should be able to show that the assemblage
nodal force vector in the following matrix equation results:
 dT ( x1 ) 
 5 5 0 0  T1  0.667 
dx 
0 0
 5 10  5 0 0 0  T   
   2  9.33

 0  5 10  5 0 0  T3   33.3 
     (6-47)
 0 0  5 10  5 0 T
  4  73.3 
0 0 0  5 10  5 T5   129.4 
    dT ( x6 
)
 0 0 0 0  5 5  T6   87.3 
 dx 

Note how the unknown derivatives (dT/dx) for the interior nodes have canceled. Physically this may
be interpreted as follows: although there is a temperature gradient and hence a heat flux at each node
and hence between two neighboring elements, these internal heat fluxes cancel. The only heat fluxes
that survive are those on the global boundary, and then only if there is actually an imposed heat flux
there. If the temperature gradients (or heat fluxes) at either or both ends were specified, the value
would simply be entered as indicated above. That case would correspond to a prescribed heat flux into
or from the ends. In any event, in the problem at hand we have prescribed temperatures at both ends,
and by using Method 1 (Method 2 may also be used), Eq. (6-47) is modified to impose T1 = 0 and
T6 = 0, which results in

1 0 0 0 0 0 T1   0.0 
0 10  5 0
 0 0 T2   9.33 
0  5 10  5 0 0 T3   33.3 
    (6-48)
0 0  5 10  5 0 T4   73.3 
0 0 0  5 10 0 T5  129.4
    
0 0 0 0 0 1 T6   0.0 

The disappearance of the (unknown) temperature gradients at nodes 1 and 6 should be noted and
thoroughly understood: in thermal problems, in general, the boundary nodes must have either a
prescribed temperature or a prescribed gradient (heat flux), convective, and/or radiative boundary
conditions. It should be noted further that the case of insulation is a special case of a specified gradient
since dT/dx = 0. This zero-gradient case is the so-called natural boundary condition. In any event,
solving Eq. (6-48) for the nodal temperatures yields

𝑇1 = 0
𝑇2 = 16.5
𝑇3 = 31.2
𝑇4 = 39.2
𝑇5 = 32.5
𝑇6 = 0

The reader will note that these values lie right on the curve for the exact solution in Fig. 6-6. The
significant point here is that two linear shape functions (or trial functions on a local or piecewise basis)
have given results that are extremely close to the exact solution.
45
40
35
Temperature, T
30
25
Exact Solution
20
Galerkin FEM
15
10
5
0
0 0.2 0.4 0.6 0.8 1
Distance along rod, x
Figure 6-6 Comparison of approximate Galerkin FEM solution with the exact solution.

6-4 APPLICATION: ONE-DIMENSIONAL HEAT TRANSFER IN A PIN FIN

Consider a circular pin fin of length Lf and varying cross-sectional area A and perimeter P as shown
in Fig. 6-7. Fins such as these are used to increase the heat transfer rate, particularly to gases. Gases
generally have low convective heat transfer coefficients. By increasing the surface area, such as in the
pin fin, we can increase the heat transfer rate significantly. The heat transfer rate is also referred to as
the heat removal rate in this text. Depending on the circumstances, fins are also commonly used in
liquids.

Figure 6-7 Circular pin fin undergoing convection to an ambient fluid.

The temperature distribution in the fin is needed for the case when the base is held at a temperature
Tb, the tip is insulated, and the fin itself convects to a fluid at a temperature Ta with a heat transfer
coefficient h. The thermal conductivity is denoted as k, and the temperature at any point along the fin
is denoted as T. The heat removal rate and fin efficiency (defined below) are to be determined also.

The governing equation may be shown to be

𝑑 𝑑𝑇
(𝑘𝐴 𝑑𝑥 ) − ℎ𝑃(𝑇 − 𝑇𝑎 ) = 0 for 0 ≤ 𝑥 ≤ 𝐿𝑓 (6-48)
𝑑𝑥

subject to the two boundary conditions


𝑇(0) = 𝑇𝑏 (6-49a)
𝑑𝑇
(𝐿𝑓 ) = 0 (6-49b)
𝑑𝑥

The intent is to obtain an approximate solution using the finite element method. In particular, the finite
element characteristics will be derived using the Galerkin method.

The Element Characteristics

The derivation begins by forming the integral of the weighted residual and setting the result to zero,
or
T d  dT  
x j N  dx  kA dx   hP(T  Ta )dx  0
xk
(6-50)

where a typical element e is assumed to connect node (at x = xj) to node k (at x = xk) as shown in Fig.
6-8. The superscript (e) on the field variable, T in this case, is no longer shown. Note that the weighting
function is taken to be the transpose of the shape function matrix, since the Galerkin method is used.
At nodes j and k, the temperatures are Tj and Tk, respectively. Note also that each node represents a
planar surface and not just a point.

Let us integrate by parts the term with the second-order derivative as follows:

𝑥 𝑑 𝑑𝑇 𝑑𝑇 𝑥𝑘 𝑥 𝑑𝐍𝑇 𝑑𝑇
∫𝑥 𝑘 𝐍 𝑇 𝑑𝑥 (𝑘𝐴 𝑑𝑥 ) 𝑑𝑥 = 𝐍 𝑇 𝑘𝐴 𝑑𝑥 | − ∫𝑥 𝑘 𝑑𝑥
𝑘𝐴 𝑑𝑥 𝑑𝑥 (6-51)
𝑗 𝑥𝑗 𝑗

The so-called integrated term cancels for all interior nodes during the assemblage step. Moreover, at
the base of the fin, the temperature is prescribed. The unknown temperature gradient at the
corresponding node is eventually eliminated when the prescribed temperature is imposed on this node.
Also, at x = Lf, we have dT(Lf)/dx = 0. Therefore, this integrated term need no longer be considered
here. Therefore, Eq. (6-50) becomes

𝑥 𝑑𝐍𝑇 𝑑𝑇 𝑥 𝑥
− ∫𝑥 𝑘 𝑘𝐴 𝑑𝑥 𝑑𝑥 − ∫𝑥 𝑘 𝐍 𝑇 ℎ𝑃𝑇𝑑𝑥 + ∫𝑥 𝑘 𝐍 𝑇 ℎ𝑃𝑇𝑎 𝑑𝑥 = 0 (6-52)
𝑗 𝑑𝑥 𝑗 𝑗

Figure 6-8 One-dimensional element with nodes j and k.


But using T = Nae and noting that only N is a function of x, we get

𝐊 𝑒 𝐚𝑒 = 𝐟 𝑒 (6-53)
where
𝐊 𝑒 = 𝐊 𝑒𝑥 + 𝐊 𝑒cv (6-54)
and
𝑥 𝑑𝐍𝑇 𝑑𝐍
𝐊 𝑒𝑥 = ∫𝑥 𝑘 𝑘𝐴 𝑑𝑥 𝑑𝑥 (6-55)
𝑗 𝑑𝑥

𝑥
𝑒
𝐾cv = ∫𝑥 𝑘 𝐍 𝑇 ℎ𝑃𝐍𝑑𝑥 (6-56)
𝑗
and
𝑥
𝐟 𝑒 = ∫𝑥 𝑘 𝐍 𝑇 ℎ𝑃𝑇𝑎 𝑑𝑥 (6-57)
𝑗
The vector of nodal unknowns ae is given by
𝑇
𝑎𝑒 = [𝑇𝑗 𝑇𝑘 ] (6-58)

and the shape function matrix N by


𝐍 = [𝑁𝑗 (𝑥 ) 𝑁𝑘 (𝑥)] (6-59)

where Nj and Nk are, in turn, given by Eqs. (6-22).

In the most general case, k, h, A, P, and Ta may all be functions of x. However, let us perform the
integrations in Eqs. (6-55) to (6-57) approximately by taking the following approach. Let us treat these
parameters as constants in any given element by using the values at the midpoint of the element where
x = 𝑥̅ and by denoting them as 𝑘̅, ℎ̅, 𝐴̅, 𝑃̅, and 𝑇̅𝑎. It can be shown that

̅̅̅̅
𝑘𝐴 1 −1
𝐊 𝑒𝑥 = [ ] (6-60)
𝐿 −1 1
̅ 𝑃̅ 𝐿
ℎ 2 1
𝐊 𝑒cv = [ ] (6-61)
6 1 2
and
̅̅̅̅̅̅̅̅̅
ℎ𝑃𝑇 𝑎 𝐿 1
𝐟𝑒 = [ ] (6-62)
2 1

where L is the element length, or L = xk - xj.

The element stiffness matrices and nodal force vectors may be determined for every element with the
help of Eqs. (6-60) to (6-62). The assemblage of these 2 x 2 matrices and 2 x 1 vectors is performed in
the usual manner to give a system of N algebraic equations in the N unknown nodal temperatures. This
assumes the fin is discretized into M elements with N nodes, where M is equal to N - l. The result is
Kªa = fa. The geometric (or prescribed temperature) boundary conditions are applied at this point to
yield Ka = f. This matrix equation may be solved for the nodal temperatures in the vector a. Therefore,
it is assumed that the nodal temperatures are known in what follows.

Heat Removal Rate

The heat removal rate can be determined by two different methods, once the nodal temperatures are
known. The first method is based on differentiation, and the second method is based on integration.
Not surprisingly, the second method is significantly more accurate, as illustrated numerically in
Example 6-3 below. In the first method, the heat removal rate QR from the fin is evaluated by
computing the heat flux from conduction at the base of the fin (where x = 0), and by multiplying this
value by the cross-sectional area A. This may be summarized as

𝑑𝑇 𝑑𝐍
𝑄𝑅 = −𝑘𝐴 𝑑𝑥 | = −𝑘𝐴 𝑑𝑥 𝐚(1) (6-63)
𝑥=0

where it is assumed that element l is at the base of the fin as shown in Fig. 6-9. It follows that
̅𝐴
𝑘
𝑄𝑅 = (𝑇1 − 𝑇2 ) (6-64)
𝐿

Figure 6-9 Portion of the discretized fin showing some elements near the base.

where it is further assumed that element 1 connects nodes 1 and 2, and where L is the element length.
More accurate results are usually obtained if the cross-sectional area at the base is used instead of the
area at the element centroid. If QR is negative, heat is conducted toward the base of the fin.

In the second method, the heat removal rate is determined by integrating the heat loss from the
exposed surfaces of the fin that undergo convection. Let us assume the tip of the fin is insulated, so
that only the fin periphery needs to be considered. Therefore, we have
𝐿
𝑄𝑅 = ∫0 𝑓 ℎ𝑃(𝑇 − 𝑇𝑎 )𝑑𝑥 (6-65)

However, it is preferred to write this as a sum of M separate integrals over each of the M elements, or

M
QR    hP(T  Ta )dx
xk
(6-66)
xj
e 1

It can be shown that if T = Nae is used, and if h, Ta, and P are evaluated at x = 𝑥̅ (i.e., at the centroid
of the element) and denoted as ℎ̅, 𝑇̅a , and 𝑃̅, then Eq. ( 6-66) becomes

M
T T 
QR   h P L j k  Ta  (6-67)
e 1  2 

where Tj and Tk are the temperatures at nodes j and k for element e. These nodal temperatures are
known from the solution of Ka = f. It is emphasized that the heat removal rate from Eq. (6-67) is more
accurate than that from Eq. (6-64).
Fin Efficiency

The fin efficiency ηf may be defined in several ways, but only one definition is given here. This
definition is based on the hypothetical condition that the entire fin is at the base temperature, because
this would result in the maximum possible heat removal rate. This corresponds to an infinite thermal
conductivity and, therefore, is not physically realizable. Nonetheless, this definition of fin efficiency
does give an indication of the relative effectiveness of the fin in increasing the heat removal rate.
Denoting this maximum rate as Qmax, we may write

𝑄𝑅
η𝑓 = 𝑄 (6-68)
max
where
𝐿
𝑄𝑚𝑎𝑥 = ∫0 𝑓 ℎ𝑃(𝑇𝑏 − 𝑇𝑎 )𝑑𝑥 (6-69)

If average values of h, P, and Ta are used for each element, Eq. (6-69) becomes

M
Qmax   h P L(Tb  Ta ) (6-70)
e 1

Again it has been assumed that the tip of the fin is insulated.

Example 6-3

Determine the temperatures, heat removal rate, and efficiency of the circular pin fin shown in Fig. 6-
10(a). The fin is made of pure copper and has a thermal conductivity k of 400 W/m-ºC. The base is
held at Tb = 85ºC and the ambient fluid temperature Ta is maintained at 25ºC. The fin length Lf is 2 cm,
and the diameter D is 0.4 cm. The convective heat transfer coefficient h is 150 W/m2-ºC and the tip at
x = Lf is insulated.

Solution

The first step in the finite element solution is discretization of the fin. Let us arbitrarily assume only
two elements (and three nodes) as shown in Fig. 6-10(b). It is convenient to summarize the node and
element data in tabular form as shown in Table 6-3. The calculations are summarized below in a form
that may be readily implemented in a computer program.

Figure 6-10 Circular pin fin (a) analyzed in Example 6-3 and (b) discretized into two elements and three nodes.
Table 6-3. Node and Element data for Example 6-3

Node number x, m

1 0.0
2 0.01
3 0.02

Element number Nodes connected


j k

1 1 2
2 2 3

Element 1: Node j is 1. Node k is 2.

xj = 0.0 m and xk =0.01 m

L= xk – xj = 0.01- 0.0 = 0.01 m


D = 0.004 m

𝜋𝐷2 𝜋(0.004)2
𝐴= = = 1.2566 x 10−5 m2
4 4

𝑃 = 𝜋𝐷 = 𝜋(0.004) = 1.2566 x 10−2 m

(1) (400)(1.2566 x 10−5 ) 1 −1


𝐊𝑥 = [ ]
0.01 −1 1

0.50265 −0.50265
=[ ] W/o C
−0.50265 0.50265

(1) (150)(1.2566 x 10−2 )(0.01) 2 1


𝐊 cv = [ ]
6 1 2

0.00628 0.00314
=[ ] W/o C
0.00314 0.00628
(1) (1) 0.50893 −0.49951]
𝐊 (1) = 𝐊 𝑥 + 𝐊 cv = [ W/o C
−0.49951 0.50893

0.50893 −0.49951 0
𝐊 𝑎 = [−0.49951 0.50893 0] W/o C
0 0 0

(150)(1.2566 x 10−2 )(0.01)(25) 1


𝐟 (1)
= [ ] = [0.23561] W
2 1 0.23561

0.23561
𝐟 𝑎 = [0.23561] W
0
Element 2: Node j is 2. Node k is 3.

xj = 0.01 m and xk =0.02 m

L= xk – xj = 0.02- 0.01 = 0.01 m

D = 0.004 m

𝐴 = 1.2566 x 10−5 m2

𝑃 = 1.2566 x 10−2 m

(2) (1)
𝐊𝑥 = 𝐊𝑥

0.50265 −0.50265]
=[ W/o C
−0.50265 0.50265
(2) (1)
𝐊 cv = 𝐊 cv
0.00628 0.00314
=[ ] W/o C
0.00314 0.00628
(2) (2)
𝐊 (2) = 𝐊 𝑥 + 𝐊 cv = 𝐊 (1)

0.50893 −0.49951]
=[ W/o C
−0.49951 0.50893

0.50893 −0.49951 0
𝐊 𝑎 = [−0.49951 1.01786 −0.49951] W/o C
0 −0.49951 0.50893

0.23561]
𝐟 (2) = 𝐟 (1) = [ W
0.23561

0.23561
𝑎
𝐟 = [0.47122] W
0.23561

Before application of the prescribed temperatures we have Kªa = fª, or

0.50893 −0.49951 0 𝑇1 0.23561


[−0.49951 1.01786 −0.49951] [𝑇2 ] = [0.47122]
0 −0.49951 0.50893 𝑇3 0.23561

Note that the assemblage stiffness or conductance matrix Kª is symmetric and banded with a half-
bandwidth of 2 [this also follows from Eq. (5-33)]. The pre-scribed temperature boundary condition at
the base of the fin will be imposed by using Method l from Sec. 5-1. The first step is to modify the
first row (since T1 is to be imposed) as shown below:

1 0 0 𝑇1 85
[−0.49951 1.01786 −0.49951] [𝑇2 ] = [0.47122]
0 −0.49951 0.50893 𝑇3 0.23561
In order to preserve symmetry, the K21 entry must be multiplied by the prescribed temperature and
transposed to the right-hand side (in the same row). Therefore, f2 must be replaced by 0.47122 - (-
0.49951) (85) or 42 .930. This is the only term that needs to be modified because K is now symmetric.
The system of equations to be solved is given by

1 0 0 𝑇1 85 85
[0 1.01786 −0.49951] [𝑇2 ] = [0.47122 − (−0.49951)(85)] = [ 42.930 ]
0 −0.49951 0.50893 𝑇3 0.23561 0.23561

Solving this system of linear, algebraic equations by the matrix inversion method yields the following
nodal temperatures:
𝑇1 = 85.0o C
𝑇2 = 81.8o C
𝑇3 = 80.8o C

The heat removal rates may now be computed as summarized below:

2
T T 
QR   h P L j k  Ta 
e 1  2 

85 + 81.8 81.8 + 80.8


= 150 𝜋 (0.004)(0.01) ( − 25) + 150 𝜋 (0.004)(0.01) ( − 25)
2 2
=1.10 + 1.06 = 2.16 W

If the alternate (less accurate) method is used, we get


(0.004)2
𝑘̅𝐴 400 𝜋 4
𝑄𝑅 = (𝑇1 − 𝑇2 ) = (85 − 81.8) = 1.60 W
𝐿 0.01

This is quite different than the result from the method based on integration. It is shown in Table 6-4
that the exact value for the heat removal rate is 2.1552 W. Clearly, the method based on integration is
significantly more accurate.

Finally, the fin efficiency is readily computed to be

𝑄𝑅 2.16 2.16
η𝑓 = = = = 0.956 or 95.6%
𝑄𝑚𝑎𝑥 150 𝜋 (0.004)(0.02)(85 − 25) 2.26

The results from Example 6-3 are compared in Table 6-4 to the exact solution given by [3]

cosh 𝜆(𝐿𝑓 − 𝑥)
𝑇(𝑥 ) = 𝑇𝑎 + (𝑇𝑏 − 𝑇𝑎 )
cosh 𝜆𝐿𝑓
and
𝑄𝑅 = 𝜆𝑘𝐴(𝑇𝑏 − 𝑇𝑎 ) tanh 𝜆𝐿𝑓
where

ℎ𝑃
𝜆=√
𝑘𝐴
Note that the results for 4 and 8 elements are also given in this table. These results were obtained by
the computer program (ABAQUS code) described below. The nodal temperatures, heat removal rates,
and efficiencies in Table 6-4 converge to the exact solution as the number of elements is increased.
Five significant digits are shown for the purpose of comparison only.

Table 6-4 Summary of Results for Example 6-3

Temperatures, oC

x, cm Exact 2 Elements 4 Elements 8 Elements

0.0 85.000 85.000 85.000 85.000


0.25 83.998 83.998
0.50 83.134 83.133 83.134
0.75 82.407 82.407
1.00 81.814 81.807 81.812 81.814
1.25 81.354 81.354
1.50 81.027 81.024 81.026
1.75 80.830 80.830
2.00 80.765 80.756 80.762 80.765

Heat removal rate, W

Case Eq. (6-64) Eq. (6-67) Fin efficiency, %*

2 Elements 1.6050 2.1618 95.6


4 Elements 1.8771 2.1569 95.4
8 Elements 2.0149 2.1557 95.3
Exact 2.1552 2.1552 95.3
*Based on Eq. (6-68) with QR from Eq. (6-67)

6-5 REMARKS

This chapter began with an introduction to some of the more important integral methods that lead
to the present-day finite element method in nonstructural applications. Only one-dimensional problems
were considered, but the basic groundwork has been laid for extension to two and three dimensions.

Four of the most popular weighted-residual methods were mentioned, which included the point
collocation method, the subdomain collocation method, the least squares method, and the Galerkin
method. These weighted-residual methods do not require any advanced mathematics beyond ordinary
calculus (i.e., variational calculus is not needed in these methods). Moreover, it was stated that the
weighted-residual methods could be used even when a classical variational principle does not exist.
The Galerkin method was cast into a form that is directly useful in the finite element method. The same
example problem was solved, approximately, by using trial functions that no longer applied globally
to the entire problem domain but rather applied locally over each element. These piecewise continuous
trial functions gave rise to two other functions that are referred to as shape functions. These particular
shape functions allow the field variable to be continuous within each element and at the ends of each
element. However, the derivatives at the ends of each element are not necessarily continuous.
Therefore, these particular shape functions are said to be C0-continuous, or to possess C0-continuity.
It was seen that accurate results could be obtained with the use of the piecewise continuous trial
functions, which are really interpolating polynomials. The implication is that the finite element method
can be used to obtain solutions to problems that possess real material discontinuities, such as composite
materials with two or more thermal conductivities.

The emphasis in this chapter has been the solution of one-dimensional, second order differential
equations. In the next chapter, as in Chapter 5, we turn to the structural analysis area where the general
finite element formulation to problems in stress analysis is presented. Two additional methods of finite
element formulations, which are more readily applied to these problems, are introduced.

6-6 THERMAL EXAMPLES USING THE ABAQUS CODE

Example 1: Steady state one-dimensional pin-fin

*HEADING
Steady State PIN-FIN *STEP
Written by E. A. Bonifaz *HEAT TRANSFER,STEADY STATE
** **
**------------------ *BOUNDARY
*PREPRINT, MODEL=NO, HISTORY=NO, ECHO=NO N1,11,,85.
*RESTART, WRITE, FREQUENCY=5 *CFLUX
*NODE, NSET=LIMITES N5,0,0.
1 **
2,0.005,0.,0. *CFILM
3,0.010,0.,0. 1,3.14159E-5,25,150
4,0.015,0.,0. 2,6.2831E-5,25,150
5,0.02,0.,0. 3,6.2831E-5,25,150
*NSET,NSET=N1 4,6.2831E-5,25,150
1 5,3.14159E-5,25,150
*NSET,NSET=N2 **
2 *OUTPUT,FIELD,FREQ=1,VARIABLE=ALL
*NSET,NSET=N3 *ELEMENT OUTPUT,ELSET=SPECIMEN
3 TEMP
*NSET,NSET=N4 *NODE OUTPUT,NSET=TODOS
4 NT
*NSET,NSET=N5 *NODE PRINT, NSET=TODOS,FREQ=1,SUMMARY=NO
5 NT
*NSET,NSET=TODOS *NODE FILE
N1,N2,N3,N4,N5 NT
** **
*ELEMENT,TYPE=DC1D2,ELSET=MASTER *OUTPUT,HISTORY,FREQ=1,VARIABLE=ALL
1,1,2 *ELEMENT OUTPUT,ELSET=SPECIMEN
*ELGEN,ELSET=SPECIMEN TEMP
1,4,1 *NODE OUTPUT,NSET=TODOS
*SOLID SECTION,MATERIAL=MAT1,ELSET=SPECIMEN NT
1.25664E-5, *NODE PRINT, NSET=TODOS,FREQ=1,SUMMARY=NO
*MATERIAL,NAME=MAT1 NT
*CONDUCTIVITY *NODE FILE
400. NT
**--------------------------------------------------------- *END STEP

ABAQUS Field Output Report


Node Label NT11
---------------------------------
1 85.
2 83.1356
3 81.8163
4 81.0296
5 80.7682
Example 2: Transient one-dimensional pin-fin

*HEADING *INITIAL CONDITIONS,TYPE=TEMPERATURE


Transient PIN-FIN TODOS,20
Written by E. A. Bonifaz *PHYSICAL CONSTANTS,ABSOLUTE ZERO=-273.16,STEFAN
** BOLTZMANN=5.66E-8
**------------------ **-----------------------------------------------------------
*PREPRINT, MODEL=NO, HISTORY=NO, ECHO=NO *STEP,INC=500
*RESTART, WRITE, FREQUENCY=5 *HEAT TRANSFER,DELTMX=150.,END=SS
*NODE, NSET=LIMITES 0.001,5.,1.E-6,,1.E-4
1 **
2,0.005,0.,0. *BOUNDARY
3,0.010,0.,0. N1,11,,85.
4,0.015,0.,0. *CFLUX
5,0.02,0.,0. N5,0,0.
*NSET,NSET=N1 **
1 *CFILM
*NSET,NSET=N2 1,3.14159E-5,25,150
2 2,6.2831E-5,25,150
*NSET,NSET=N3 3,6.2831E-5,25,150
3 4,6.2831E-5,25,150
*NSET,NSET=N4 5,3.14159E-5,25,150
4 **
*NSET,NSET=N5 *OUTPUT,FIELD,FREQ=1,VARIABLE=ALL
5 *ELEMENT OUTPUT,ELSET=SPECIMEN
*NSET,NSET=TODOS TEMP
N1,N2,N3,N4,N5 *NODE OUTPUT,NSET=TODOS
** NT
*ELEMENT,TYPE=DC1D2,ELSET=MASTER *NODE PRINT, NSET=TODOS,FREQ=1,SUMMARY=NO
1,1,2 NT
*ELGEN,ELSET=SPECIMEN *NODE FILE
1,4,1 NT
** **
*SOLID SECTION,MATERIAL=MAT1,ELSET=SPECIMEN *OUTPUT,HISTORY,FREQ=1,VARIABLE=ALL
1.25664E-5, *ELEMENT OUTPUT,ELSET=SPECIMEN
*MATERIAL,NAME=MAT1 TEMP
*SPECIFIC HEAT *NODE OUTPUT,NSET=TODOS
449.944 NT
*DENSITY *NODE PRINT, NSET=TODOS,FREQ=1,SUMMARY=NO
8890. NT
*CONDUCTIVITY *NODE FILE
400. NT
** *END STEP
Example 3: Thermo-mechanical model in a one-dimensional pin-fin

*HEADING 552,0,871
stresstransient PIN-FIN 345,0,982
Written by E. A. Bonifaz *EXPANSION,ZERO=20
**------------------ 13.3E-6
*PREPRINT, MODEL=NO, HISTORY=NO, ECHO=NO **
*RESTART, WRITE, FREQUENCY=5 *INITIAL CONDITIONS,TYPE=TEMPERATURE
*NODE, NSET=LIMITES TODOS,20
1 **
2,0.005,0.,0. **
3,0.010,0.,0. **----------------------------------------------------------------------
4,0.015,0.,0. *STEP,INC=500
5,0.02,0.,0. *STATIC
*NSET,NSET=N1 0.001,5.,1.E-6,,1.E-4
1 *TEMPERATURE,FILE=transient-pinfin,BSTEP=1,BINC=1,ESTEP=1,EINC=61
*NSET,NSET=N2 *BOUNDARY
2 N1,1
*NSET,NSET=N3 N5,1
3 **
*NSET,NSET=N4 **-------------------------------------------------------------------
4 **
*NSET,NSET=N5 *NODE PRINT, NSET=TODOS,FREQ=5,SUMMARY=NO
5 COORD,U
*NSET,NSET=TODOS *EL PRINT,ELSET=SPECIMEN,FREQ=5,SUMMARY=NO,POSITION=CENTROIDAL
N1,N2,N3,N4,N5 E,S
*ELEMENT,TYPE=T2D2,ELSET=MASTER **
1,1,2 *OUTPUT,FIELD,FREQ=1,VARIABLE=ALL
*ELGEN,ELSET=SPECIMEN *ELEMENT OUTPUT,ELSET=SPECIMEN,POSITION=CENTROIDAL
1,4,1 *NODE OUTPUT,NSET=TODOS
*SOLID SECTION,MATERIAL=MAT1,ELSET=SPECIMEN U,RF
1.25664E-5, **
*MATERIAL,NAME=MAT1 *OUTPUT,HISTORY,FREQ=1
*ELASTIC *ELEMENT OUTPUT,ELSET=SPECIMEN
204E3,0.3 E,S
*PLASTIC *NODE OUTPUT,NSET=TODOS
952.,0. U,RF
952.2,0,21 **
952.2,0,600 *NODE FILE
910.8,0,649 U,RF
793.5,0,760 *END STEP
Example 4: Transient three-dimensional plate-fin

*HEADING 1,48541,60
Three dimensional transient plate-fin *SURFACE,NAME=IZQUIERDA
Written by: E. A. Bonifaz IZQUIERDA,S6
** *NSET,NSET=IZQUIERDA,ELSET=IZQUIERDA
**------------------ **
*PREPRINT, MODEL=NO, HISTORY=NO, ECHO=NO *ELSET,ELSET=DERECHA,GENERATE
*FILE FORMAT,ZERO INCREMENT 60,48600,60
*RESTART, WRITE, FREQUENCY=5 *SURFACE,NAME=DERECHA
*NODE, NSET=LIMITES DERECHA,S4
1 *NSET,NSET=DERECHA,ELSET=DERECHA
61,120.E-3,0.,0. **
901,0.,18.E-3,0. *ELGEN,ELSET=ABAJO
961,120.E-3,18.E-3,0. 1,60,1,1,90,1000,540
90001,0.,0.,180.E-3 *SURFACE,NAME=ABAJO
90061,120.E-3,0.,180.E-3 ABAJO,S3
90901,0.,18.E-3,180.E-3 *NSET,NSET=ABAJO,ELSET=ABAJO
90961,120.E-3,18.E-3,180.E-3 **
*NSET,NSET=N1 *ELGEN,ELSET=ARRIBA
1 481,60,1,1,90,1000,540
*NSET,NSET=N61 *SURFACE,NAME=ARRIBA
61 ARRIBA,S5
*NSET,NSET=N901 *NSET,NSET=ARRIBA,ELSET=ARRIBA
901 **
*NSET,NSET=N961 **----------------------------
961 *ELSET,ELSET=TEST
*NSET,NSET=N90001 24210,24211,24270,24271,24750,24751,24810,24811
90001 *NSET,NSET=TEST,ELSET=TEST
*NSET,NSET=N90061 **
90061 **============================================
*NSET,NSET=N90901 **
90901 *SOLID SECTION,MATERIAL=MAT1,ELSET=SPECIMEN
*NSET,NSET=N90961 1.,
90961 *MATERIAL,NAME=MAT1
**------------------------------ *SPECIFIC HEAT
*NFILL,NSET=BOTO 418.4,21
N1,N61,60,1 460.24,93
*NFILL,NSET=TOPO 502.08,204
N901,N961,60,1 523,316
*NFILL,NSET=BOTF 543.92,427
N90001,N90061,60,1 565,538
*NFILL,NSET=TOPF 585.76,649
N90901,N90961,60,1 627.6,760
**---------------------------- 669.44,871
*NFILL,NSET=BACK 711.28,982
BOTO,TOPO,9,100 711.28,1093
*NFILL,NSET=FRONT 660.46,1260
BOTF,TOPF,9,100 705,1340
*NFILL,NSET=TODOS 763,2780
BACK,FRONT,90,1000 763.42,2980
** 763.46,3000
*ELEMENT,TYPE=DC3D8,ELSET=MASTER 763.46,5000
1,1,2,102,101,1001,1002,1102,1101 *DENSITY
*ELGEN,ELSET=SPECIMEN 8110
1,60,1,1,9,100,60,90,1000,540 *CONDUCTIVITY
** 14,21
*ELSET,ELSET=ATRAS,GENERATE 15,93
1,540,1 11.8,204
*SURFACE,NAME=ATRAS 13.7,316
ATRAS,S1 15.57,427
*NSET,NSET=ATRAS,ELSET=ATRAS 17.7,538
** 19.7,649
*ELSET,ELSET=ADELANTE,GENERATE 21.5,760
48061,48600,1 23.3,871
*SURFACE,NAME=ADELANTE 25.4,982
ADELANTE,S2 27.2,1093
*NSET,NSET=ADELANTE,ELSET=ADELANTE 28.06,1100
** 29.0,1120
*ELSET,ELSET=IZQUIERDA,GENERATE 30.11,1140
31.34,1200 **----------------------------------------------
32.16,1240 **
32.8,1300 *NODE PRINT, NSET=TEST,FREQ=5,SUMMARY=NO
31.32,1320 NT
30.05,1328 *EL PRINT,ELSET=TEST,FREQ=5,SUMMARY=NO,POSITION=CENTROIDAL
28.97,1340 TEMP,COORD
28.97,1360 **
28.97,5000 *OUTPUT,FIELD,FREQ=1,VARIABLE=ALL
** *ELEMENT OUTPUT,ELSET=TEST,POSITION=CENTROIDAL
*INITIAL CONDITIONS,TYPE=TEMPERATURE *NODE OUTPUT,NSET=TEST
TODOS,20 NT
*PHYSICAL CONSTANTS,ABSOLUTE ZERO=-273.16,STEFAN **
BOLTZMANN=5.66E-8 *OUTPUT,HISTORY,FREQ=1
** *ELEMENT OUTPUT,ELSET=TEST
**---------------------------------------------- TEMP
*STEP,INC=500 *NODE OUTPUT,NSET=TEST
*HEAT TRANSFER,DELTMX=150.,END=SS NT
0.001,1200.,1.E-6,,1.E-4 **
** *NODE FILE
*BOUNDARY NT
ATRAS,11,,85. **
*CFLUX *END STEP
ADELANTE,0,0.

Results of the Simulation:

REFERENCES

1. Stassa, F. L., Applied Finite Element Analysis for Engineers. HRW Series in Mechanical
Engineering. CBS Publishing -Japan Ltd. 1985.
2. Becker, E. B., G. F. Carey, and J. T. Oden, Finite Elements: An Introduction, vol. 1, Prentice-Hall,
Englewood Cliffs, N.J., 1981, pp. 6-9.
3. Incropera, F. P., and D. P. DeWitt, Fundamentals of Heat Transfer, Wiley, New York, 1981, pp.
105- 117.

PROBLEMS

Refer to Stassa’s book [1]

You might also like