Anda di halaman 1dari 48

Lecture

Note for
MEE 511: An Introduction to the
FINITE ELEMENT METHOD



















Pattaramon Jongpradist

Department of Mechanical Engineering
King Mongkuts University of Technology Thonburi



Preface


This lecture note present an introduction and basic aspects of Finite Element
Method (FEM) in the context of solid mechanics. The book is for the graduate course
MEE511 Finite Element Method offered in the Department of Mechanical Engineering.
of the King Mongkuts University of Technology Thonburi. The objective of the book is
to balance the theoretical knowledge on mathematical concept of FEM as well as the
physical interpretation through applications of the method to real world examples.
Computer implementation is also introduced so that student understand what goes
behind the scenes of a commercial product and able to use the commercial product
effectively with confidence. The ultimate goal is that students feel comfortable with
the FEM fields such that they can employ the method to any of the advanced problems
they might encounter during their professional life.

ii

Chapter 1
Fundamental of Finite Element Method

1.1 Introduction

The Finite Element Method (FEM) is a numerical procedure to obtain solutions of
the equations that govern the physical phenomena found in nature. These physical
behaviors can usually be expressed in differential or integral form so-called governing
equation. Thus, the FEM is a numerical technique for solving partial differential or
integral equations.
Mathematical models of the FEM have been formulated for the many physical
phenomena in engineering systems. Common physical problems solved using the
standard FEM include:
Mechanics of solids and structures
Heat transfer
Fluid flow analysis
Electro-magnetism
Acoustics
Coupled physics problems (e.g., fluid-structure interaction)
Others
In the analysis of structures, the FEM is a method for computing the
displacements, stresses and strains in a structure under a set of loads. This is the main
objective of this book. Having a clear understanding of the basic concepts of FEM will
enable engineers to use a general-purpose finite element software, such as
Hyperworks, effectively. Throughout this text, emphasis is placed on basic theory and
background concepts for finite element analysis (FEA) and methods for which to verify
the obtained FE results.

1.2 Computational Modeling using FEM

Analytical methods search for the universal mathematical expressions
representing the general and exact or closed-form solution of a problem governed.
Unfortunately, exact solutions are only possible for a few particular cases, which
frequently represent coarse simplifications of reality with assumptions and relaxed
constraints. Practical problems usually involve complicated domains and conditions.
Numerical or computational methods such as FEM aim to providing an approximate
solution in the form of a set of numbers to the mathematical equations governing a
problem. Most numerical methods transform the mathematical expressions into a set
of algebraic equations depending on a finite set of parameters. The number of
equations is usually rather large for most real-world applications, and requires the
computational power of the digital computer.
In FEM, a finite element can be visualized as a small portion or subdomain of a
continuum. The word finite distinguishes such an element from the infinitesimal
elements of differential calculus. The geometry of the continuum is considered to be
formed by the assembly of a collection of non-overlapping subdomains with simple
geometry. The continuum has an infinite number of degrees-of-freedom (DOF), while
the discretized model has a finite number of DOF. This is the origin of the name, finite

element method. Elements are connected and defined by using nodes. Triangles and
quadrilaterals in two dimensions or tetrahedra and hexahedra in three dimensions are
typically chosen to represent the elements. The continuum domain is said to be
discretized by a mesh of finite elements. Some examples of element discretization in
structural models are shown in Fig. 1.1.
The space variation of the problem parameters (i.e. the displacements of the
structure) is expressed within each element by means of a polynomial expansion. Since
the exact analytical solution is complex and generally unknown, the FEM only provides
an approximation to the exact solution. Fig. 1.2 shows an example of finite element
approximation for a one-dimensional rod AB. A continuous function is approximated
using piecewise linear functions in each element.
This book focuses on the formulation of finite element equations for the
mechanics of solids and structures. However, the application of the FEM to all other
physical problems utilized similar concepts.


Fig. 1.1. Discretization of different solids and structures with finite elements [1].

Unknown&discrete&values&of&
u(x)% Unknown&func*on&of&eld& eld&variable&at&nodes"
variable,&u(x)"

uh3(x)"
uh2(x)"
&u h
1(x)"

Approximated&func*ons&

x"
A B
elements" nodes"


Fig. 1.2. Finite element approximation for a one-dimensional case (adapted from [7])

The governing system of equations for solids can be derived by using equilibrium
equations in the case of static analysis or Dalemberts principle for dynamic problems.
The partial differential equations governed the system are of strong forms which
require a strong continuity on the dependent field variables (in this case, the
displacements u, v and w). Obtaining the exact solution of the strong form of the
governing equation is possible for simple and regular geometry and boundary
conditions and thus usually very difficult for practical engineering problems.
A weak form of the system equations is usually created using energy principles [2]
or weighted residual methods [3]. The energy principle is a special form of the
variational principle, which is particularly suited for problems of the mechanics of
solids and structures. The weighted residual method is more general for solving all
kinds of partial differential equations.
The weak form is often an integral form and requires a weaker continuity on the
field variables. Weak form formulations usually produce a set of discretized system
equations which give much more accurate results especially for problems with
complex geometry. Hamiltons energy principle will be introduced for FEM formulation
in the following section. The approach works out the dynamic system equations. The
dynamic terms can simply be dropped out when static system equations are of interest.


1.3 Basic Steps in the Finite Element Method

The basic steps in any finite element analysis consist of the following:

a. Preprocessing Phase

1. Create and discretize the problem domain into finite elements; that is,
subdivide the problem into nodes and elements.
2. Assume a shape function to represent the physical behavior of an element;
that is, a continuous function is assumed to represent the approximate
behavior (solution) of an element.
3. Develop equations for an element

4. Assemble the elements to present the entire problem. Construct the global
stiffness matrix
5. Apply boundary conditions, initial conditions, and loading.

b. Solution Phase

6. Solve a set of linear or nonlinear algebraic equations simultaneously to obtain
nodal results, such as displacement values at different nodes or temperature
values at different nodes in a heat transfer problem.

c. Postprocessing Phase

7. Obtain other important information, such as, principal stresses, heat fluxes,
etc.

1.4 FEM Procedure

This section briefly summarizes the standard FEM procedure and illustrates the
steps involved in direct formulation.

a. Preprocessing Phase

1.4.1 Domain Discretization

The procedure to divide the solid body into elements is called meshing or
discretization. The purpose of discretization is to make it easier in assuming the
pattern of the displacement field. Figure 1.3 shows an example of a mesh for a two-
dimensional solid. An element is formed by connecting its nodes. All the elements
together form the entire domain of the problem without any gap or overlapping. The
domain can consist of different types of elements with different numbers of nodes as
long as they are compatible (no gap and overlapping) on the boundaries between
different elements.

!!
!! !!
!! !!
!! !!
Domain,!! !! e! !! !!
!! !!
!!
!! !! Element!
!!!!!! !!
!!
Boundary,!! Node!

Inter1element!ux! !!
!!
!!
!! !! Domain,!h!
!! e!
!!

!! !!



Fig. 1.3 Example of a mesh with elements and nodes numbered.


The density of the mesh depends upon the accuracy requirement of the analysis
and the available computation resources. Finer mesh generally yields more accurate
results but increase the computational cost. Thus, finer mesh is used in the areas of
large displacement gradient or where the accuracy is critical to the analysis.
To illustrate the FEM steps, let us consider a bar with variable cross section as
shown in Fig. 1.4. The bar is fixed at one end and carries the load P at the other end.
The width of the bar varies from w1 at the top to w2 at the bottom. Its thickness is t, its
length is L and the bars modulus of elasticity is E. The weight of the bar is assumed
negligible and we are interested in approximating the deflection of the bar at various
points along its length.
Fig. 1.5 is an example of meshing of the bar into 5 nodes and 4 elements. Each
element is modeled by an average area of cross sections. Note that the accuracy of FE
results can be increased by generating a model with additional nodes and elements.

1.4.2 Assume displacement Interpolation of an element

Consider an element with nd nodes at coordinates " = " , " , " where i = 1, 2, ..,
nd. The displacement approximation functions are assumed in the form of a linear
combination of nd linearly independent basis functions pi(x), i.e. for displacement u in
x-direction,




w1"

y" L"



w2"



P"

Fig. 1.4 A bar under axial loading


1"
l1" u1"
1"
element"1"
A1"
2"
2" u2"
y" A
l2" element"2"
L" 2"
3" u3"
3"
l3" element"3"
A3"
4" 4"
l4" u4" element"4"
A4"
5" 5"
u5"

P" P" P"

Fig. 1.5 Discretizing the bar into elements and nodes


*
= 2 ()" = 5 (1.1)

where ai is the coefficient for each basis function pi. A complete basis function of order
k is built based on the well-known Pascals pyramid in Fig. 1.6. The basis function
terms should be selected from the constant term to higher orders symmetrically from
the Pascal pyramid.
The coefficients are determined by enforcing the value of * in Eq. (1.2) to be
equal to the nodal displacements at the nd nodes, . Thus, at node i we have,

= 5 i = 1, 2, 3, ..., nd (1.2)

In the matrix form,

= (1.3)

where is the vector that includes the values of the displacement component at all
the nd nodes in the element = {d1 d2 ... ;< }T. The matrix P is called the moment
matrix given by

5 (= ) = (= ) > (= ) ;< (= )
5
( > ) ( ) > ( > ) ;< ( > )
= = = > (1.4)

5 ( ;< ) = ( ;< ) > ( ;< ) ;< ( ;< )

Assuming the inverse of P exists, we then have

= B= (1.5)


Fig. 1.6. Pascal pyramid of monomials

Substituting Eq. (1.5) into Eq. (1.2), we therefore obtain



* = 2 ()" = (1.6)

where is the matrix of shape functions Ni(x) defined by

= 5 () B= (1.7)

Note that in selecting the interpolation functions for displacements, the following
guidelines should be considered.

1. The common type of interpolation functions are polynomials.


2. The interpolation function should be continuous within the element to prevent
openings, overlaps and jumps of displacement value.
3. The interpolation function should provide interelement continuity for all
degrees of freedom along boundaries. In other words, the function is conforming
or compatible.
The symbol Cm describes the order of continuity, where superscript m indicates
the degree of derivative that is interelement continuous.
4. The interpolation function should allow for rigid-body displacement within the
element, i.e., the function must capable of yielding a constant value. Also, a state
of constant strain must also be allowed.
The polynomial satisfying this guideline is said to be complete function, i.e., the
lower order terms cannot be omitted. Completeness is necessary condition for
convergence to the exact answer as the number of elements is increased (Fig.
1.7)

0

-0.001
One element
Two elements
-0.002
Four elements
Displacement

Eight elements
-0.003
Exact solution

-0.004

-0.005

-0.006
0 10 20 30 40 50 60

Axial coordinate (along the bar)

Fig. 1.7 Comparison of exact and finite element solutions for axial displacement
of a rod with uniform cross-section and linearly distributed axial load

Properties of shape function are as follow:


1) If the complete order of polynomial is k, the shape function is said to
possess Ck-1 consistency (compatibility). This also means that the
displacement field can be exactly reproduced using the shape functions as
long as the given field function is included in the basis functions
(reproduction property).
2) Shape functions are linearly independent. This is because basis functions
are of linear independence and shape functions are equivalent to the basis
functions in function space.
3) Shape functions possess delta function properties. The shape function Ni
should be unit at its node i and vanishes at the remote nodes , or
1 = , = 1, 2, . . , N
" H = "H =
0 , , = 1, 2, . . , N
4) Shape functions are partitions of unity:
;
"2= " = 1

For the bar example (Fig. 1.5) along y-axis, a two-node element is used to
model a uniform cross-section bar with the length l as shown in Fig. 1.8. From Pascals
triangle, a complete first order basis function, () = , is applied. From Eq. (1.2)
the approximated displacement within the element is

=
* = 5 = 1 = = + > ()
>

Substitute the coordinates and displacements at nodes (node 1: = = 0, = = , and


5 ( ) 1 = 1 0
> = 0, = > ) to obtain = 5 = = = , we get
( > ) 1 > 1

= 1 0 =
= : = ()
> 1 >

fi#

i#
ui#
y# kequivalent#
l#

A#
i+1#
ui+1#

fi+1#

P#
Fig. 1.8 A linear spring element representing a solid member of uniform cross section

The vector a is calculated by = B= as

= 1 0 B= = = 0 =
> = 1 > V 1 1 > ()
=

The displacement approximation within the element is * = 5 = , where


the shape function N is calculated from Eq. (1.7) as

= 0 Z Z
= 5 B= = 1 = [1 V ] ()
V 1 1 V

Therefore, the linear interpolation function of displacement within the element is


obtained as a function of the displacement at nodes as

Z Z =
* = [1 V V
] ()
>

* =

From the obtained interpolation functions, let consider if they are appropriate
according to the guidelines. First, the shape function is linear function that is
continuous within the bar as shown in Fig. 1.9. Next, since the degrees of freedom for
the bar are the axial displacements at nodes (= and > ), it is required that the element
is C0 continuous. This means that only the displacement field itself is continuous across
the common node 2 in Fig. 1.10, i.e., interelement compatibility is satisfied for
displacement field. Its first derivative (strain) in the element 1 and 2 are constant
values, i.e., (=) = (> -= )/ and (>) = (c -> )/ . These two constant strains are not
necessarily equal, that is, they are not interelement continuous. For rigid body
displacment, let consider the case where the displacement within the element is a
constant value or = . Therefore,
= = = = >
From Eq. (1.6) and Eq. (e), we have
= = = + > > = (= + > )=
Therefore, it requires that
= + > = 1
This leads to one of the shape function properties that the shape functions must add to
unity at every point within the element so that u will yield a constant value when a
rigid body displacement occurs.

1 1

> () =
= () = 1

y
= >

Fig. 1.9 Distribution of shape functions for 1D element

c >
y> () = +
> = c > >Z c > cZ
y= () = =Z +
> = > = >Z




= 1 > 2 c y

Fig. 1.10 Interelement compatibility

1.4.3 Formation of FE Equations in Local Coordinate System

To derive generalized FE equations, the discretized dynamic system equations
are first considered and will later be simplified to static equilibrium equations.
Hamiltons principle states Of all the admissible time histories of displacement the most
accurate solution makes the Lagrangian functional a minimum.
An admissible displacement must satisfy the following conditions: compatibility
equations, the essential or the kinematic boundary conditions, and the conditions at
initial (t1) and final time (t2).
Mathematically, Hamiltons principle states:
gh
gi
= 0 (1.8)
The Lagrangian functional L is obtained by

= + n (1.9)
where T is the kinetic energy, is the potential energy (in our cases, elastic strain
energy) and n is the work done by the external forces. The kinetic energy is defined
by
=
= 5 (1.10)
> q

Substituting the interpolation of the nodes, Eq. (1.1) into Eq. (1.10), the
expression for kinetic energy in an element is
= =
= 5 5 = 5 t (1.11)
> qs > qs

Let y = qs
5 be the mass matrix of the element, Eq. (1.11) can be
rewritten as
=
= 5 y (1.12)
>

The strain energy in the elastic solids can be expressed as


= =
= 5 = 5 (1.13)
> q > q

10

where and are the strain and stress matrices, c is a matrix of material constants
which gives constitutive relations = . For the strain matrix B relating strain-
displacement equation = (see Section 2.3)
= =
= 5 5 = 5 5 (1.14)
> qs > qs

By denoting the element stiffness matrix y

y = qs
5 (1.15)
Eq. (1.19) can be rewritten as
=
= 5 (1.16)
>

Finally, the work done by external forces can be calculated by

n = qs
5 5 + s
5 5
5
= qs
5 + 5 5 = 5 + 5 = 5 y
s
(1.17)

and are the nodal body and surface forces acting on the nodes of the
elements and the total nodal force vector = + .

Substituting Eqs. (1.12), (1.16) and (1.17) into Lagrangian function L Eq. (1.9),
we have
gh = 5 =
gi >
5 + 5 = 0 (1.18)
>

The variation and integration operators are interchangeable. Also, the variation
and differentiation with time are interchangeable. Integrating the first term by parts,
we obtain
gh gh gh
5 y = 5 5
gi gi gi
gh
= gi
5 (1.19)

Substituting Eq. (1.19) into Eq. (1.18) leads to


gh
gi
5 + = 0 (1.20)

The variation of the displacements is arbitrary, the condition to satisfy Eq. (1.20) is

+ = (1.21)
Eq. (1.21) is the dynamic FEM equation for an element, while and are the
stiffness and mass matrices for the element, and is the element force vector of the
total external forces on the nodes of the element.
In static analysis, the matrix = . The element FE equation is therefore,

= (1.22)

11


For the bar example, the relation of the applied force P and the elongation d
within the elastic region is

= ()

The bar is represented by a model consisting of four elastic spring (elements) in series,
and the behavior of an element with nodes i and i+1 is modeled by an equivalent linear
spring (Fig. 1.6) according to the equation,

- " + "=
= y "= " = "= " = " ()
2 "=
where the equivalent element stiffness is given by

(" + "= )
y = ()
2

The transmitted forces at nodes i and i+1 calculated from static equilibrium
condition of the element are

" = y " "=
()
"= = y "= "

Equation () can be expressed in a matrix form by

" y y "
= "= ()
"= y y

1.4.4 Assembly of Global FE Equation

The FE Equations for all the individual elements can be assembled together to
form the global FE system equation

+ = (1.23)

where K and M are the global stiffness and mass matrix, D is the vector of all the
displacements at all the nodes in the entire problem domain, and F is a vector of all the
equivalent nodal force vectors. For static analysis, the FE equations reduce to

= (1.24)

The assembly process is done by adding up the contributions from all the
elements connected at a node. To assemble all element equations to form the global
system equations, a coordinate transformation has to be performed for each element.
The coordinate transformation matrix T gives the relationship between the
displacement vector based on the local coordinate system and the displacement
vector for the same element based on the global coordinate system.

12

= T (1.25)

Also, the force vectors can be transformed between the local ( ) and global
coordinate systems ( ) as

= T (1.26)

Substitution of Eq. (1.25) and (1.26) into Eq. (1.24) leads to element equation
based on the global coordinate system

+ = (1.27)

where

= 5 (1.28)

and = 5 (1.29)

For the bar example, the stiffness matrix for element (1) is given by

= =
(=) = ()
= =

The transformation matrix T1 for element (1) relating the element nodal displacement
= (Fig. 1.6) to the displacement in global coordinate system (Fig. 1.5) can be
written by using Eq. (1.25) as,
=
>
= 1 0 0 0 0 c
> = 0 1 0 0 0 ()

The position of (=) in the global stiffness matrix is given by Eq. (1.28), (=) =
5 (=) ,

= = 0 0 0 =
= = 0 0 0 >
(=) = 0 0 0 0 0 c ()
0 0 0 0 0
0 0 0 0 0

Note that, for this simple 1D spring example, the matrix can also be written by
direct formulation without the use of T1. Similarly, the stiffness matrices for element 2
to 4 are

> > c c
(>) = , (c) = , () = ()
> > c c

and their positions in the global stiffness matrix are as follow:

13

0 0 0 0 0 = 0 0 0 0 0 =
0 > > 0 0 > 0 0 0 0 0 >
(>) = 0 > > 0 0 c , (c) = 0 0 c c 0 c
0 0 0 0 0 0 0 c c 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 c ()
0 0 0
0 0 0

The global stiffness matrix obtained by assembling (adding together) element stiffness
matrices according to their positions in the global matrix:

= (=) + (>) + (c) + ()

= = 0 0 0
= = + > > 0 0
= 0 > > + c c 0 ()
0 0 c c +
0 0 0

1.4.5 Apply boundary conditions and loads

Next, the structures have to be constrained against rigid body movement. The
constraints can be imposed by removing the rows and columns in Eq. (1.29)
corresponding to the constrained nodal displacements. If the constraints are sufficient,
the remaining stiffness matrix K will be of full rank and will be of Symmetric Positive
Definite (SPD) property.

For the bar example, the bar is fixed at the top (= = 0). The external load P is
applied only at node 5, while the external load at nodes 2, 3 and 4 are zero. The
external load at node 1 is equal to the reaction force = . We therefore obtain a set of
linear equations:

= = 0 0 0 0 =
= = + > > 0 0 > 0
0 > > + c c 0 c = 0 ()
0 0 c c + 0
0 0 0

Removing the first row and column corresponding to the constraint = = 0, we
obtain the FE equation that can be solved for solution as

= + > > 0 0 > 0
> > + c c 0 c 0
= ()
0 c c + 0
0 0

14

b. Solution Phase

1.4.6 Solving the Global FE Equation

By solving the global FE equation, displacements at the nodes can be obtained.
The strain and stress in any element can then be retrieved by the strain-displacement
relations and the stress-strain relations, as will be explained in Chapter 2.

For other classes of problems, the FE procedure is basically the same. Table 1.1.
shows examples of variables for engineering problems governed by second order
differential equation of the form:
N N
+ = 0 (1.30)
N N

Table 1.1 Variables of engineering problems governed by second order PDE [2]

Field&of&study Primary& Secondary&
Problem&data
variable& variable&
u a c f Q
Heat&transfer Temperature Thermal&conductance Surface&convection Heat&generation Heat
T kA AP f Q
Flow&through& Fluid&head Permeability 0 Infiltraction Point&source
porous&medium f Q
Flow&through&pipes Pressure Pipe&resistance 0 0 Point&source
P 1/R Q
Flow&of&viscous& Velocity Viscosity 0 Pressure&gradient Shear&stress
fluids v ,dP/dx xz
Elastic&cables Displacement Tension 0 Transverse&force Point&force
u T f P
Elastic&bars Displacement Axial&stiffness 0 Axial&force Point&force
u EA f P
Torsion&of&bars Angle&of&twist Shear&stiffness 0 0 Torque
GJ T
Electrostatics Electrical&potential Dielectic&constant 0 Charge&density Electric&flux
E

Table 1.2 Properties of the elements for the bar example



Element Nodes Aavg (mm2) l (mm) E (N/mm2) keq (N/mm)
1 1 2 112.5 50 70 x 103 157500
2 2 3 97.5 50 70 x 103 136500
3 3 4 82.5 50 70 x 103 115500
4 4 5 67.5 50 70 x 103 94500

For the bar example, let us assume that E = 70000 MPa (aluminum), = = 40
mm, > = 20 mm, t = 3 mm, L = 200 mm and P = 8000 N. The element stiffness
coefficient for each spring are computed from Eq. (h) as shown in Table 1.2.
Assembling the elemental matrixes leads to the global stiffness matrix:

15

157.5 157.5 0 0 0
157.5 157.5 + 136.5 136.5 0 0
= 10c 0 136.5 136.5 + 115.5 115.5 0 ()
0 0 115.5 115.5 + 94.5 94.5
0 0 0 94.5 94.5

Applying the boundary condition = = 0, the first row and the first column
corresponding to the constraint is removed. The following 4 x 4 matrix is to be solved:

294 136.5 0 0 > 0


136.5 252 115.5 0 c 0
10c = ()
0 115.5 210 94.5 0
0 0 94.5 94.5 8000

The displacement solution is = = 0, > = 0.0508 mm, c = 0.109 mm, = 0.179 mm,
and = 0.263 mm.

c. Postprocessing Phase

1.4.7 Obtain other important information.

Displacement function within the element can be determined from local coordinate
system by Eq. (e). At y = 120 mm in global coordinate, the displacement value can be
obtained from local equation of element 3. Substitute y = 20 mm, = = c = 0.154 mm
and > = = 0.213 mm,
Z Z > >
* = 1 = + > = 1 0.109 + 0.179 = 0.137 ()
V V

The average normal stresses in each element can be determined from



= = " ()
- "=

=
The average stress in element 1 is = = 0.0508 0 = 71.1 MPa.

The average stresses in elements 2, 3 and 4 are > = 82.1 MPa, c = 97.0 MPa, and
= 119 MPa.

The reaction force at node 1 can be computed from the first row of Eq. (q),

= = = > = = 157500 0.0508 0 = 8000 N

The obtained reaction R1 can be shown to satisfy the static equilibrium of the entire
bar.

1.5 Verification and Validation of FEM Results

Confidence in using the obtained FEM results is quantified by using the
verification and validation methods. Verification is the process of determining that a
computational model accurately represents the underlying structural model and its
solution. This is made by comparing the numerical results for simple benchmark
problems with analytical solutions or solutions from more accurate numerical

16

methods. Validation is the assessment of the accuracy of the computational models


comparing to the numerical results from experimental data using scale models or
actual structures. In other words, verification is performed to evaluate if the
mathematic equations are being solved correctly, whereas validation is to ensure that
the right equations are being solved. Figure 1.11 shows a scheme of the verification
and validation steps from ASME [4].


























Fig. 1.11 Scheme of the verification and validation processes in the FEM [4].

17

Chapter 2
Stress-Strain Analysis


Stress analysis is a major step in the mechanical design process. The main
considerations in design are [5] (i) the stress at every point should be below a certain
limit for the material; (ii) the deflection should not exceed the maximum allowable fro
proper functioning of the system; (iii) the structure should be stable; and (iv) the
structure or machine element should not fail by fatigue.
This chapter introduces basic concepts in mechanics of solids under statics
condition including the concept of stress and strain, the relationships between stresses
and strains, and the relationships between displacements and strains. This book deals
only with problems of very small deformation. Thus, the materials show linear elastic
properties. The material of interest is also limited to isotropic material.

2.1 Stress

2.1.1 Surface Traction
Consider a solid subjected to external forces and in static equilibrium as shown in
Fig. 2.1. To determine the state of stress at a point K in the interior of the solid, the
body is cut into two halves by passing an imaginary plane through Q. The unit vector
normal to the plane is denoted by n. The internal forces acting on the cut surface are
such that the portion of the body is in equilibrium. Surface traction is defined as the
internal force per unit area or the force intensity acting on the cut plane. The traction at
Q is measured by the force acting over a small area that contains point Q. The
surface traction () acting at the point Q is defined as

() = lim (2.1)


The traction T is a vector which can be resolved into components and write it as
() = + Z + (2.2)
The superscript (n) denotes the normal n of the plane that this surface traction is
defined. Note that at the same point , the traction vector T would be different on a
different plane passing through Q. In general, the directions of T and n are not
necessarily parallel to each other. The surface traction T can be decomposed into two
components. i.e., the component normal to the plane or parallel to n called the normal
stress (; ), and the other parallel to the plane or perpendicular to n called the shear
stress (; ). f4#
f3# f3#
F"

Q" n"
f2#
f2# A#

f5#

f1# f1#

Fig. 2.1 Surface traction acting on a plane at a point

20

2.1.2 Rectangular Stress Components



Since the surface traction at a point varies depending on the direction of the
normal to the plane, an infinite number of traction vectors () and corresponding
normal and shear stresses for a given state of stress at a point can be obtained. The
state of stress at a point can be completely defined by specifying the traction vectors on
three orthogonal planes passing through the point. These planes can be taken as the
planes normal to the x, y and z-axes.

Consider the yz-plane, which is normal to the x-axis. The traction vector can be
written as
() () ()
() = + Z + (2.3)

() () ()
The traction is the normal stress, Z and are the shear stresses in the y- and
z-directions. These components can be denoted by , Z and , or

() = + Z + (2.4)

Similarly, the other stress components can be obtained by passing two more planes,
normal to y and z-axes,
= Z + ZZ + Z
(2.5)
= + Z +

The stress components can be depicted using a cube as shown in Fig. 2.2. The sign
convention for the subscript is that the first letter represents the surface on which the
stress is acting, and the second letter represents the direction of the stress. The positive
directions of the stress components are as shown in Fig. 2.2. By equilibriums of forces
and moments of the infinitesimal element, it can be proved that only six independent
stress components are required to identify the state of stress at a point.
Consider an arbitrary plane of normal n through the point represented as


= + Z + = Z (2.6)


z


Z


Z
ZZ
. Z y
Z


x

Fig. 2.2 Stress components in Cartesian coordinate system

21

z

C



Z
B y
Q

A
x

Fig. 2.3 Surface traction on an arbitrary plane with normal n

The traction () can be computed by using equilibrium of the tetrahedron QABC as
shown in Fig. 2.3, where the limit of the distance h from point Q to each vertex
approaches zero. Let the areas of the triangles ABC, QAB, QBC, and QAC be A, Anz, Anx,
Any, respectively. Force balance in x-direction yields

(;)
= Z Z = 0 (2.7)

Dividing Eq. (2.7) by A, we obtain
(;)
= + Z Z + (2.8)

Similarly, force balance in the y- and z-directions yields
(;)
Z = Z + ZZ Z + Z
(;)
(2.9)
= + Z Z +

Therefore, the surface traction acting on the surface whose normal is n can be
determined from the stress components as

() = (2.10)
Z
where = Z ZZ Z is a stress matrix that completely characterizes the state of
Z
stress at a given point.

Example 2.1 The stress at a point P is given below. The direction cosines of the normal
n to a plane passing through P have the ratio nx: ny: nz = 4: 3: 12. Determine (a) the
traction vector () , (b) the normal stress ; , (c) the shear stress ; .

26 13 26
= 13 13 0 MPa
26 0 39

Solution: The dimension of the specify vector is 4> + 3> + 12> = 13
=
The unit normal vector n is =
=c
4 3 12 5
The surface traction is obtained by Eq. (2.10) as

22


26 13 26 4 19
=
() = = 13 13 0 3 = 1
=c
26 0 39 12 28
The normal stress is calculated by
Bc =>
; = () = 19 + 1 + 28 = 53.38 MPa
=c =c =c
>
() = 19> + 1> + 28> = 1146

; = () > 53.38> = 21.26 MPa



2.1.3 Principal Stresses

The normal and shear stresses acting on a plane passing through a given point in
a solid change, as the orientation of the plane is changed. Therefore, there must be a
plane on which the normal stress is the maximum and another plane on which the
shear stress attains a maximum. These planes have significance in predicting the failure
of the material at that point.
The principal stresses are the normal stresses that attain the extremum (maximum
or minimum) values. The principal stresses can be proved to act on three mutually
perpendicular planes on which the shear stresses vanish. On these planes, the traction
() is parallel to the normal vector n or () = ; . Thus, from Eq. (2.10)

= ; (2.11)

Eq. (2.11) is the eigenvalue problem and can be arranged as

; = 0

; Z 0 (2.12)
Z ZZ ; Z = 0
Z ; 0

Physically meaningful solution exists if and only if the determinant of the coefficient
matrix is zero, i.e.,
; Z
Z ZZ ; Z = 0 (2.13)
Z ;

Expanding this determinant, the cubic equation in ; is obtained as

;c = ;> + > ; c = 0 (2.14)
where
= = + ZZ +

Z ZZ Z
> = ZZ + Z +
Z

> > >
= ZZ + ZZ + Z Z

23


> > >
c = ZZ + 2Z Z Z ZZ Z

In the above equation, = , > , and c are the three invariants of the stress matrix ,
which are independent of the coordinate system. The three roots of Eq. (2.14)
correspond to the three principal stresses.
Once the principal stresses have been computed, the stresses are substituted into
Eq. (2.12) one at a time to obtain the three linear simultaneous equations with three
unknown , Z , and . The equations are however not independent since the matrix
is singular (the determinant is zero). Therefore, another equation based on the fact that
n is a unit vector must be satisfied;

> = > + Z> + > = 1 (2.15)

The obtained normal vector n1, n2, n3 are the principal directions on which the three
principal stresses act.

2.2 Equations for Three-Dimensional Solids

2.2.1 Stress-Strain Relations

The constitutive equation gives the relationship between the stress and the strain
following Hookes law. For three-dimensional elastic solid, the generalized Hookes law
can be given as
= (2.16)

where the stress tensor, = ZZ Z Z
5
the strain tensor, 5 = ZZ Z Z
and c is the elasticity matrix. For isotropic material, c can be written as

1 0 0 0
1 0 0 0
1 0 0 0
=
(=B>)(=) 0 0 0 0 0
0 0 0 0 0
0 0 0 0 0
in which E, n and G are Youngs modulus, Poissons ratio, and the shear modulus of the
material, respectively. The relationship between these three constants is G = E/2(1 + n).

Strain is the change of displacement per unit length. The strain components can
be obtained from the derivatives of the displacements as

= , ZZ = , = (2.17)
Z Z

Z = + , = + , Z = +
Z Z

where u, v and w are the displacement components in the x, y and z directions,


respectively.

24

The strain-displacement relationships can be written in the matrix form as

= (2.18)

where u is the displacement vector = and L is a matrix of partial


differential operators
/ 0 0
0 / 0
0 0 /
=
0 / /
/ 0 /
/ / 0

2.2.2 Equilibrium Equations

Consider an infinitely small block of solid as shown in Fig. 2.4. The static
equilibrium of forces in the x direction gives

+ + Z + Z Z
+ + = 0 (2.19)

where = , Z = , = and is the x component of
Z
the external body force f applied at the center of the element.

Z + Z
+ Z + Z
+
ZZ + ZZ
+
dz
Z + Z
z Z + Z dy

y dx
x

Fig. 2.4 Stresses on an infinitesimal 3D element.

Eq. (2.19) becomes the equilibrium equation of forces in x-direction as



+ + + = 0 (2.20)
Z

Similarly, the equilibrium equations of forces in y and z directions are



+ + + Z = 0
Z

(2.21)
+ + + = 0
Z

25

Two types of boundary conditions in elasticity problems are displacement


(essential) and force (natural) boundary conditions. The displacement boundary
conditions are used to describe the support or constraints which the displacement is
zero (homogeneous boundary condition). A homogeneous force condition implies a free
surface.

2.3 Equations for Two-Dimensional Solids

Three-dimensional problems can be simplified as a two-dimensional solid when
the variables are independent of one coordinate axis (usually the z-axis). Two common
types of 2D problems are plane stress and plane strain.
In the plane stress condition (Fig 2.5), the thickness of the solid in the z direction
is very small compared with dimensions in x and y directions. External forces are
applied only in the x-y plane and stress components in the z-direction ( , Z , ) are
all zero. In this case, the strains and Z are zero, but will not be zero due to the
effect of Poissons ratio.































Fig. 2.5 Example of plane stress problems. Displacement field and loads acting on the
middle plane section [1].

26


The plane strain condition (Fig. 2.6) occurs when the thickness in the z-direction
of the solid is very large compared with the dimensions in the x and y directions.
External forces are applied evenly along the z axis, the movement in the z direction at
any point is constrained. The strains components in the z direction ( , Z , ) are all
zero. In the case, the stresses and Z are zero, but will not be zero.






























Fig. 2.6 Examples of plane strain problems. Displacement field and loads acting on a
transverse section [1].

2.3.1 Stress-Strain Relations

The system equation for 2D problems can be obtained by omitting terms related
to the z direction in the equations for 3D problems. The stress and strain components
are 5 = ZZ Z and 5 = ZZ Z .

For plane stress conditions of = Z = = 0, the constitutive equation Eq.
(2.16) for isotropic material is obtained as

27

1 0
1 0
ZZ = ZZ (Plane stress) (2.22)
=B h =
Z 0 0 (1 ) Z
>

For plane strain problems where = Z = = 0, the stress-strain relation is

1 0
1 0
ZZ = ZZ (Plane strain) (2.23)
(=)(=B>) =
Z 0 0 Z
>

The strain-displacement relationships are

=
/ 0
ZZ = 0
/ (2.24)
Z
/ /

Example 2.2 The displacement field for the thin cantilever beam of length L subjected to
a point load P at one end as shown in the figure is
y, v P

x, u


L

> c
, =
2 6
>
> c
, =
2 2 6
where I is the moment of inertia about bending axis. Determine the stress field.
Solution
For a thin beam, a plane stress condition is assumed. The strain components are

= =


ZZ = =

> > > >
Z = + = + = 0
2 2 2 2

Substituting into Eq. (2.22), the stress field is obtained as
>
= >
=
1

28


ZZ = + = 0
=B h

Z = 0

The equilibrium equations for 2D solids can be obtained by eliminating the terms
related to the z coordinate from Eq. (2.20) (2.21):

+ + = 0
Z

(2.25)
+ + Z = 0
Z

2.4 Equations for Truss Members

A truss member in a truss structure is a solid where dimension in one direction is
much larger than in the other two directions. The connection of the truss structure is
assumed as a pin joint if the centerlines of the members are concurrent at the joint. In
analysis, all external forces are assumed to be applied at the pin connections. The
internal force is then in the axial direction. Therefore, a truss member is actually a one-
dimensional solid.
The equations for 1D solids can be obtained by further omitting the stress related
to the y direction from the 2D case. The only stress in a truss member is or , and
its corresponding strain is or .

Hookes law for 1D solids without the terms in y direction and Poisson effect can
be written in the form
= (2.26)

The strain-displacement relationship is given by

= (2.27)


The static equilibrium equation in 1D solids is

+ = 0 (2.28)


Substitute Eq. (2.26) and (2.27) into Eq. (2.28) to obtain the equilibrium equation
in terms of displacement for elastic and homogenous trusses
h
+ = 0 (2.29)
h

For truss members or bars of constant cross-sectional area A, Eq. (2.29) becomes
h
+ = 0 (2.30)
h

where F = f A is the external force applied in the axial direction of the bar.

29

2.5 Equations for Beams



Beams and trusses are different such that the forces applied on beams are
transversal or the force direction is perpendicular to the axis of the beam. The beam
experiences bending and its deflection is in y direction as a function of x.
Thin-beam theory is referred to as the Euler-Bernoulli beam theory. It assumes
the plane cross-sections, which are normal to the undeformed centroidal axis, remain
plane after bending and remain normal to the deformed axis as shown in Fig. 2.7. This
assumption means that the shear stress Z is negligible. Secondly, the axial
displacement u at a distance y from the centroidal axis is

! = (2.31)
where is the rotation in the x-y plane that can be obtained from the deflection in the y-
direction of the centroidal axis of the beam, Z , as

= (2.32)

The relationship between the normal strain and the deflection can be given by

h
= = (2.33)
h


z









Fig. 2.7 Euler-Bernoulli assumption for thin beams

From Hookes law, the normal stress is


h
= = (2.34)
h

The moments resulting from linearly distributed normal stress on the cross-section of
the beam can be calculated by integrating over the area of the cross-section
h h
= = > = (2.35)
h h

where I is the second moment of inertia of the cross-section with respect to the y-axis.

30


A B
C C
x Dx
L
qdx

V
M M+dM

Dx
V+dV

Fig. 2.8 Simply supported beam subject to a distributed load

For a beam subject to distributed load q as shown Fig. 2.8, a small element CC of length
dx is cut out (Fig. 2.8b). For force equilibrium in the y direction,

+ + = 0
Nq
or = () (2.36)
N

Consider the moment equilibrium of the beam element with respect to the right surface
of the element,
N
+ = 0 (2.37)
>

Neglecting the second order small term (dx)2,


N
= (2.38)
N

Substituting Eq. (2.36) and (2.38) into Eq. (2.35), we obtain the differential equation of
a beam with constant EI as

= (2.39)

The beam equation will be explained further in Chapter 3.


31

Chapter 3
FEM for One-Dimensional Elements


A spring is one of the simplest structural members designed to take only axial
forces along the orientation of the springs. A structure made from a system of springs is
a one-dimensional structure. A truss is a straight bar with the dimensions of the cross-
section much smaller than that in the axial direction. Therefore, it also takes only the
axial loads and deforms only in its axial direction. However, a truss system can be in
two-dimensional planes and in three-dimensional space.
This chapter develops finite element equations for such spring and truss
members via the direct stiffness method. Then, the principle of minimum potential
energy is applied to derive the finite element equations for truss structure.

3.1 FE Equations for Springs: Direct Stiffness Approach

Consider a linear spring with nodes 1 and 2 at the ends of the element as shown in
Fig. 3.1. Let the forces acting on node 1 and 2 are F1 and F2, respectively, acting in the
direction of the spring associated with local axis x. The displacements at node 1 and 2,
u1 and u2, are called degrees of freedom at each node. Positive direction of the forces and
displacements at each node are taken in the positive x as shown in the figure.

The force on the spring is obtained from the multiplication of the spring stiffness
and the spring deformation. Thus, the force F1 on the spring at node 1 equal to k(u1 u2)
where k is the spring stiffness (or spring constant). Similarly, the force F2 at node 2 is
k(u2 u1). In other words, F1 = F2.

We can rewrite the above equations in matrix form as:

= =
= (3.1)
> >

or = (3.1a)

where = = > 5 is the nodal force vector, = = > 5 is the nodal displacement

vector, and the matrix = is the element stiffness matrix.

k
f1 1 2 f2 x

u1 u2

Fig. 3.1 Linear spring element with positive nodal displacement and force conventions.

For a linear spring, Eq. (3.1) can be generalized as:



= => =
= == (3.2)
> >= >> >

54

where kij is the force on ith node induced by a unit displacement in the jth node.

3.2 FE Equations for 2D Planar Truss: Direct Stiffness Approach

Truss member is a solid where dimension in one direction is much larger than in
the other two directions. The connection of the truss structure is assumed as a pin joint.
All external forces are assumed to be applied at the pin connections. Thus, a truss
member is a two-force member with its internal force only in the axial direction (no
shear, no moment). Therefore, a truss member is a 1D solid.

Consider a structure consisting of a number of trusses or bar members. Each of
the members can be considered as a truss element of uniform cross-section bounded by
two nodes. A truss element along a local x-axis is shown in Fig. 3.2. The length of the
element is y . The applied force at node 1 and 2 are = y
and y> , and the nodal
displacements at node 1 and 2 are = y
and y> . The stiffness of the truss element is y .
The prime symbol () indicates the local coordinate of the truss. Since a truss member is
an axial member, only forces and displacements along the x-axis exist. The superscript
e denotes the elemental force and displacement values.

The internal force pe is then computed as
y
y = y> = = = y y (3.3)

where y is the cross-sectional area of the truss element and y is the normal stress of
y
the truss section subjected to the external loads = and y> .

For 1D elastic deformations, the stress-strain relationship from Hookes law is


y = y y where y is Youngs modulus of the element. The strain y in 1D truss is thus
calculated from

y
y> =
y
= (3.4)
y
Substituting Hookes law equation and Eq. (3.4) into Eq. (3.3), we obtain
y y y y y
y>
= =
= y > = (3.5)

The element stiffness equation in local coordinate of a truss member is then written as
y y
= y y 1 1 =
= y y (3.6)
y> 1 1 >

or = (3.6a)


2


1
Fig. 3.2 Basic truss element with positive nodal forces and displacement.

55

Similarly, for a linear torsional element, the element stiffness equation in local
coordinate can be expressed as
y
y
= y y 1 1 =
= y (3.6)
y> y 1 1 >
y
where where = = y> 5 is the nodal torsion vector indicated internal torsion
y y 5
at nodes about x-axis, = = > is the vector of nodal twisting angles about x-
axis, y is materials shear modulus and y is polar moment of inertia of the members
cross section.

y
>Z y
, >Z


y y y
y y 2 > , > y
=Z , =Z
x

x

1 y , y
= =

Fig. 3.3 Nodal forces and displacements in the global x-y coordinates of a truss element.

To analyze a truss element in the global coordinates x and y in Fig. 3.3, both
components of the displacement have to be accounted for. For a general truss element
oriented at the angle q measured counter-clockwise from x to x, four nodal
y y y y
displacement components must be considered, i.e., = , =Z , > , >Z .

Consider a transformation of vectors in two dimensions as shown in Fig. 3.4. A


displacement vector u can be written in x-y and x-y coordinate as
= + Z ! = + Z ! (3.7)

where and ! are unit vectors in x and y directions, and and ! are unit vectors in x
and y directions. The unit vectors and ! can be given in terms of the unit vectors and
! as
= !
(3.8)
! = + !

Substitution of Eq. (3.8) into Eq. (3.7), we obtain the relations between the x-y
coordinates and the x-y coordinates as
! + Z + ! = + Z ! (3.9)

In a matrix form,
cos sin
= (3.10)
Z sin cos Z

56

y
y
x
!
q
sin !
x





Fig. 3.4 Vector transformations.

Therefore, the displacement transformation from x, y-axes to displacements


along x and y axes are expressed as
y y
= cos sin 0 0 =
y y
=Z sin cos 0 0 =Z
= y (3.11)
y> 0 0 cos sin >
y>Z 0 0 sin cos y
>Z

or = (3.11a)
y y
= =
y y
=Z =Z
where = and
= y are the nodal displacement vector in the local
y> >
y
y>Z >Z
x-y and global x-y coordinate, respectively.
cos sin 0 0
= sin cos
0 0 is the transformation matrix.
0 0 cos sin
0 0 sin cos

It can be easily proven that is orthogonal matrix or = where is the identity
matrix of 44.

Similarly, we can transform the nodal force vector in the local and global
coordinates as
= (3.12)

From the element stiffness matrix equation in local coordinate (3.6a), pre-
multiply to both sides of the equation. We then obtain the element stiffness matrix
equation based on the global coordinate system:

= (3.13)

or = (3.14)

where the element stiffness matrix in the global coordinate, , is

57

= (3.15)

cos > sin cos cos > sin cos


=
s s sin cos sin> sin cos sin>
#s cos > sin cos cos > sin cos
sin cos sin> sin cos sin>

The stiffness matrix has 22 symmetric substructure. Therefore, the reverse
numbering of nodes (1 and 2) results in the same element stiffness matrix. The element
stiffness matrix in Eq. (3.15) is singular (does not have an inverse). Physical
interpretation is that if the nodal displacements are specified, the element forces can be
uniquely determined by using Eq. (3.14). However, if forces acting on the elements are
given, the nodal displacement cannot be uniquely determined because one can always
translate the element by adding a rigid body displacement without affecting the forces
acting on it. Therefore, it is always necessary to remove the rigid body motion by fixing
some displacement at nodes.

The assembly process in this case is the same as for spring structures. After the
element stiffness equations for all elements are obtained. The equations are rewritten
to include all nodal displacements according to the global degrees of freedom instead of
local ones. Consider the free-body diagram of a typical node i connected to element e
and e+1 as shown in Fig. 3.5. The acting forces include the external force Fi , and
reactions to the element forces "y and "y= . The internal forces are acting in
direction because they are the reaction to the forces acting on the element. The
equilibrium condition is such that

" "y "y= = 0 (3.16)

or " = "y "y= (3.16a)

Eq. (3.16) shows that the external force acting on a node is equal to the sum of all
the internal forces acting on all y elements connected to the node or
"s
" = y
y2= "
(3.17)

The element equations are then be assembled to a single global stiffness matrix
equation. Afterwards, the prescribed essential (displacement) and natural (force)
boundary conditions are substituted and all unknown nodal displacements can be
determined from the equation

= (3.18)

where K is the global stiffness matrix, d are the unknown displacements, and F are the
known external forces applied to nodes. Note that the global stiffness component Kij is
contributed by elements between nodes i and j, whereas element that share node i
contributes to the global stiffness component Kii. The global stiffness matrix is
always a positive definite matrix that has an inverse. It is square symmetric and the

58

diagonal elements of K are positive, i.e., "" > 0. Thus, the displacement d can be solved
uniquely for a given set of nodal forces F.

After all nodal displacement are obtained, the internal element forces can be
determined from Eq. (3.14). The strain is computed from Eq. (3.4) as
y
= =
y = 1 1 (3.19)
#s y>

"
"
(y)
"
(y=)

Element e Element e+1
Node i
Fig. 3.5 Force equilibrium at node i
Therefore, the stress is obtained from
y
=
y = 1 1 (3.20)
#s y>

Example 3.1: Consider the truss in the figure below. Determine the deflection of each
joint. All members are made from wood with a modulus of elasticity of E = 13.1 GPa and
a cross-sectional area of 50 cm2. Also, calculate average stresses in each member.

Solution:

a. Preprocessing phase

1. Domain Discretization

Each truss member is considered an element and each joint is a node. Table 3.1
shows the properties of each element.

59

Table 3.1 Properties for each element

Element Node i Node j q L (m) k (N/m)

(1) 1 2 0 1 65.510&
(2) 2 3 135 2 46.310&

(3) 3 4 0 1 65.510&

(4) 2 4 90 1 65.510&
(5) 2 5 45 2 46.310&
(6) 4 5 0 1 65.510&


2. Assume displacement interpolation of an element

The element is modeled as a spring with an equivalent stiffness of y = #
as shown in
=
Table 3.1. An example of stiffness constant for element 1 is

(13.1109 )(50104 )
(=)
= = 65.5106 N/m
1.0
3. Formulation of FE equation for elements

From Eq. (3.15), the stiffness matrices are

cos > 0 sin 0 cos 0 cos > 0 sin 0 cos 0


>
() = 65.510& sin 0 cos 0 sin 0 sin 0 cos 0 sin> 0
> >
cos 0 sin 0 cos 0 cos 0 sin 0 cos 0
>
sin 0 cos 0 sin 0 sin 0 cos 0 sin> 0
1 0 1 0 =
0 0 0 0 =Z
() = 65.510&
1 0 1 0 >
0 0 0 0 >Z

Similarly,
0.5 0.5 0.5 0.5 > 1 0 1 0 c
0.5 0.5 0.5 0.5 >Z & 0 0 0 0 cZ
() = 46.310& ()
c , = 65.510 1
0.5 0.5 0.5 0.5 0 1 0
0.5 0.5 0.5 0.5 cZ 0 0 0 0 Z

0 0 0 0 > 0.5 0.5 0.5 0.5 >


0 1 0 1 >Z & 0.5 0.5 0.5 0.5 >Z
() = 65.510& ()
, = 46.310 0.5
0 0 0 0 0.5 0.5 0.5
0 1 0 1 Z 0.5 0.5 0.5 0.5 Z

1 0 1 0
0 0 0 0 Z
() = 65.510&
1 0 1 0
0 0 0 0 Z

60

4. Assemble elements

The global stiffness matrix is obtained by assembling the element stiffness matrix as
65.5 0 65.5 0 0 0
0 0 0 0 0 0
65.5 0 65.5 + 23.2 + 23.2 23.2 + 23.2 23.2 23.2
0 0 23.2 23.2 65.5 + 23.2 + 23.2 23.2 23.2
0 0 23.2 23.2 65.5 + 23.2 23.2
= 10&
0 0 23.2 23.2 23.2 23.2
0 0 0 0 65.5 0
0 0 0 65.5 0 0
0 0 23.2 23.2 0 0
0 0 23.2 23.2 0 0
0 0 0 0 0 =
0 0 0 0 0 =Z
23.2 0 0 23.2 23.2 >
23.2 0 65.5 23.2 23.2 >Z
23.2 65.5 0 0 0 c

23.2 0 0 0 0 cZ
0 65.5 + 65.5 0 65.5 0
0 0 65.5 0 0 Z
0 65.5 0 65.5 + 23.2 23.2
0 0 0 23.2 23.2 Z

Simplifying, we obtain
65.5 0 65.5 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0
65.5 0 111.8 0 23.2 23.2 0 0 23.2 23.2
0 0 0 111.8 23.2 23.2 0 65.5 23.2 23.2
0 0 23.2 23.2 88.7 23.2 65.5 0 0 0
= 10&
0 0 23.2 23.2 23.2 23.2 0 0 0 0
0 0 0 0 65.5 0 131 0 65.5 0
0 0 0 65.5 0 0 0 65.5 0 0
0 0 23.2 23.2 0 0 65.5 0 88.7 23.2
0 0 23.2 23.2 0 0 0 0 23.2 23.2

5. Apply the boundary conditions and loads

The displacement (essential) boundary conditions are: nodes 1 and 3 are fixed.
Therefore, = = 0, =Z = 0, c = 0, and cZ = 0. The external loads (natural boundary
conditions) at nodes 4 and 5 are such that Z = 1200 N, and Z = 1200 N. A set of linear
equations from Eq. (3.18) to be solved is:

65.5 0 65.5 0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0 0 0
65.5 0 111.8 0 23.2 23.2 0 0 23.2 23.2 > 0
0 0 0 111.8 23.2 23.2 0 65.5 23.2 23.2 >Z 0
& 0 0 23.2 23.2 88.7 23.2 65.5 0 0 0 0 0
10 =
0 0 23.2 23.2 23.2 23.2 0 0 0 0 0 0
0 0 0 0 65.5 0 131 0 65.5 0 0
0 0 0 65.5 0 0 0 65.5 0 0 Z 1200
0 0 23.2 23.2 0 0 65.5 0 88.7 23.2 0
0 0 23.2 23.2 0 0 0 0 23.2 23.2 Z 1200

61


The rows and columns corresponding to the prescribed displacements, i.e., the first,
second, fifth, and sixth rows and columns, can be eliminated. Thus, only a 66 matrix is
needed to be solved as

111.8 0 0 0 23.2 23.2 > 0
0 111.8 0 65.5 23.2 23.2 >Z 0
0 0 131 0 65.5 0 0
10& Z =
0 65.5 0 65.5 0 0 1200
23.2 23.2 65.5 0 88.7 23.2 0
23.2 23.2 0 0 23.2 23.2 Z 1200


b. Solution phase

6. Solve a system of algebraic equations simultaneously



Solving the above equation, we get

> 0.0553
>Z 0.1594
0.0183
Z = mm
0.1777
0.0366
Z 0.3031
c. Postprocessing phase

7. Obtain other information



Reaction forces
The reaction forces can be computed from
=

=
65.5 0 65.5 0 0 0 0 0 0 0
=Z
0 0 0 0 0 0 0 0 0 0
> 65.5 0 111.8 0 23.2 23.2 0 0 23.223.2
>Z 0 0 0 111.8 23.2 23.2 0 65.5 23.223.2
c & 0 0 23.2 23.2 88.7 23.2 65.5 0 0 0
cZ = 10 0 0 23.2 23.2 23.2 23.2

0 0 0 0

0 0 0 0 65.5 0 131 0 65.5 0
Z 0 0 0 65.5 0 0 0 65.5 0 0
0 0 23.2 23.2 0 0 65.5 0 88.7 23.2
Z 0 0 23.2 23.2 0 0 0 0 23.2 23.2
0 0
0 0
0.0553 0
0.1594 0
0 0
10Bc
0 0
0.0183 0
0.1777 1200
0.0366 0
0.3031 1200

62

Therefore, the reaction forces are


= 3622
=Z 0
> 0
>Z 0
c 3616
cZ = 2416 N
0
Z 0
0
Z 0

Internal forces and normal stresses

First, the local displacements in terms of the global displacements are obtained from Eq.
(3.11). As an example, element 5 connecting nodes 2 and 5 is considered,

> cos 45 sin 45 0 0 0.0553


>Z sin 45 cos 45 0 0 0.1594
= 103
0 0 cos 45 sin 45 0.0366
Z 0 0 sin 45 cos 45 0.3031

The member internal forces can be obtained from the element equation Eq. (3.14);

> 0.5 0.5 0.5 0.5 0.0553


>Z 0.5 0.5 0.5 0.5 0.1594
= 46.310& 103
0.5 0.5 0.5 0.5 0.0366
Z 0.5 0.5 0.5 0.5 0.3031

> 1197
>Z 1197
= N (compression)
1197
Z 1197

The normal stress in element 5 is calculated from Eq. (3.20) to be compressive stress of
339 kPa.

Similarly, the internal forces and stresses can be obtained for other elements.

To verify the results, we can use the computed reactions forces and the external forces
to check for static equilibrium or check whether the sum of the forces at each node is
zero.

3.3 FE Equations for 3D Space Truss: Direct Stiffness Approach

For a space truss in three-dimensional problems, the element stiffness matrix
equation based on the local coordinate system is as given in Eq. (3.6). However,

63

displacements and forces in the truss elements are in the x, y and z-axes as shown in
Fig. 3.4. The global coordinates of node 1 and 2 are at (= , = , = ) and (> , > , > ).
The length of the element is
y = (> = )> + (> = )> + (> = )> (3.21)

The displacement transformation of a is defined as
y
=
y
=Z
y y
= y y y 0 0 0 =
= y (3.22)
y> 0 0 0 y y y >
y
>Z
y
>

or = (3.22a)
2 1 Zh BZi 2 1
where y =

, = , y =

are the direction cosines of the axial axis of
#s
the element. The matrix is the transformation matrix. The transformation matrix
also applies to the force vectors between the local and global coordinate systems:

= (3.23)

y
> y
, > y
>Z y
, >Z


y y
y y y y 2 > , >
= , = =Z , =Z z, /
y,


y y
1 = , = x,

Fig. 3.6 A two-node truss element in x-y-z coordinate system.

The stiffness matrix for 3D truss based on global coordinate system is therefore,

= (3.24)


y > y y y y y > y y y y
y y y > y y y y y > y y


= y y y y y > y y y y y >
y > y y y y y > y y y y
y y y > y y y y y > y y
y y y y y > y y y y y >

The procedure for the assembly of individual elemental matrices for a space-truss
member (including applying boundary conditions, loads and solving for displacements)
is exactly the same as in the case of a two-dimensional truss

64


3.4 Principle of Minimum Potential Energy

An alternative approach to derive the element equation is Variational
Formulation. The method is more general than direct stiffness method and more
suitable for complicated elements. For solid mechanics problems, a functional of the
total potential energy called the principle of minimum potential energy is used. The
principle of minimum potential energy is stated that:
Of all the geometrically possible and compatible displacements of a body,
corresponding to the satisfaction of stable equilibrium, the one that minimizes the total
potential energy is the exact solution.
The total strain energy stored in a body with given displacements ( y ) is equal to
the potential energy for these given displacements (0 y ) and the work done by external
loads on these displacements ( y ).

0 y = y + y (3.25)

For a finite element 1D rod, the elastic strain energy is


= y y
y = 1 >
y (3.26)
where y , y is the stress and strain in the element.

The term y is the potential energy of the external loads (Fig. 3.7) is given by
y y y y
y = q
i
(= = + > > ) (3.27)

where the first term on the right hand side of Eq. (3.28) represent the potential energy
of body forces , typically from the self-weight (force per unit volume) moving
through displacement function u. The second term is the surface loading or traction ,
typically from distributed loading acting along the surface of the element (force per
unit surface area) moving through displacement . The third and fourth terms are
y y y
nodal concentrated forces, = and > , moving through nodal displacements, =
y
and > , at nodes 1 and 2, respectively.


(y) (y)
= >

Fig. 3.7 General forces acting on a one-dimensional bar

The negative sign indicates that the potential energy of external forces is lost
when the work is done by the external forces.

The axial displacement function u can be written in terms of nodal displacement
as
= (3.28)

65

where N is the shape function matrix and d is the nodal displacement vector or

= 1 (3.29)


and = = (3.30)
>

The displacement function over the surface that the distributed surface traction acts
is described by the shape function Ns evaluated over that surface and the nodal
displacement d as
= (3.31)

Then, using the strain-displacement relationship = , we can write the strain
as
1 1 =
= (3.32)
>

or = (3.32a)

N 1 1
where B = = (3.33)
N

The axial stress-strain relationship is given by


= (3.34)
where for 1D case, D = E is elastic modulus of the material. Therefore,
= (3.35)

From Eq. (3.25), the potential energy is

0 y = (3.36)
2 1 q i

where both and are in general column matrices. Thus, it is necessary to place the
transpose on for proper matrix multiplication. The vector P represents the
concentrated nodal loads P = = > . Substitution of Eqs. (3.28), (3.31), (3.32) and
(3.35) into Eq. (3.36), we obtain

0 y = (3.37)
2 1 q i

It can be seen that 0 y is a function of nodal displacement vector d. However, B, D and


d are not functions of x. Therefore, integrating Eq. (3.37) with respect to x yields


0 y = (3.37)
2

where = q
+ i
+ (3.38)

66

To minimize the total potential energy, the potential energy 0 y must be


differentiated with respect to all degrees of freedom.
(89 s )
= (3.39)
:

First, the term y is explicitly defined as


1
y
1 1 =
y = y
= y> 1
[] y
2

>

> y y >
y = y
= 2= > + y> (3.40)
2 2

The term is explicitly written as


y
= = y
= +y> y> (3.41)

Applying this principle of minimum potential energy in Eq. (3.39) to one truss element
by taking partial derivatives of Eq. (3.37) with respect to the nodal displacements (local
y y
coordinates) = and > .

( ) s s
s = 1 2 = 1 (3.42a)
< i #s

( ) s s
s = 2 1 = 2 (3.42b)
< h #s

Equation (3.42) can be rearranged to be of the same form as Eq. (3.6).

In general, the principle of minimum potential energy of an element takes the


form:
=
0 y = (3.43)
>

where the element stiffness matrix = .

After assembly, we obtain

=
0 = (3.44)
>

The minimization gives us the solution, = or

67

Example 3.2:
A bar of length L is subjected to a linearly distributed axial loading that varies from zero
at node 1 to a maximum at node 2. Determine the energy equivalent nodal loads.



Solution: Use Eq. (3.31) and the shape function from Eq. (3.29)
y
=
= y =
> i

# 1
=


>
c

= 2 3
c

3 0
>
= 6>

3
Note that the total load in the area under the load distribution is given by
1 > y y
= = = = + >
2 2


3.5 Higher Order Elements

3.5.1 Quadratic Elements

The accuracy of the FE results can be increased by the number of linear


elements or by using higher order interpolation functions. When a quadratic
interpolation function is used instead of a linear function, three nodes are required to
define an element. The nodes are at the ends and the mid-length as shown in Fig. 3.8.
The displacement distribution for a typical element is represented by
, = = + > + c > (3.45)

Substitute the coordinates for nodes 1, 2, 3 (= , > , c ) and the nodal displacements
= , > , c

68

= 1 = => =
> = 1 > >> > (3.46)
c 1 c c> c

Displacement distribution can be written in terms of the nodal displacement as
=
, = = = > c > (3.47)
c
where the shape functions are
2
= = ( > )( c )
>
>
> = ( = )( c ) (3.47)
#h
4
c = ( = )( > )
>

u x u x

d1 d1

d3
d3 d4
d2 d2

x x
1 3 2 1 3 4 2
L/2 L/3 L/3 L/3
L L

(a) (b)
Fig. 3.8 Displacement distribution in an element (a) Quadratic approximation
(b) Cubic approximation

3.5.2 Cubic Elements

The cubic interpolation function for displacement field in one-dimensional problem is
, = = + > + c > + c (3.48)
Displacement distribution can be written in terms of the nodal displacement as
=
>
, = = = > c (3.49)
c

where, in this case, the shape functions are
9
= = ( > )( c )( )
2c
9
> = c ( = )( c )( )
2
69

27
c = = > (3.50)
2c
27
= c ( = )( > )( c )
2
Note that when the order of interpolating function increases, it is more
convenient to employ Lagrange interpolation functions instead of using the above
approach.
The Lagrange method calculates the shape functions in terms of product of linear
functions. For a cubic function, the shape function at node i is calculated as product of
three linear functions that their product produces a value of unity at that node and a
value of zero at other nodes. The Lagrange polynomial formula to calculate the shape
function for node 1 of an ( 1) order can be written as
( > )( c ) ( ; )
= = (3.51)
(= > )(= c ) (= ; )
The shape functions for other nodes can be obtained similarly.

Example 3.3:
Using Lagrange interpolation functions, generate the quadratic shape functions directly
from the Lagrange polynomial formula Eq. (3.51).

Solution:
For quadratic shape functions (order 2), n = 3.
( > )( c ) ( > )( c ) 2
= = = = > ( > )( c )
(= > )(= c )
( )()
2
( = )( c ) ( = )( c ) 2
> = = = > ( = )( c )
(> = )(> c )
()( )
2
( = )( > ) ( = )( c ) 4
c = = = > ( = )( > )
(c = )(c > )
( )( )
2 2
The results are the same as Eq. (3.47).

70

Anda mungkin juga menyukai