Anda di halaman 1dari 131

THE DEVELOPMENT OF A FINITE ELEMENT TOOL FOR THE CALCULATION

OF BEAM CROSS SECTION PROPERTIES

by

BRYAN DALE MIXON

Presented to the Faculty of the Graduate School of

The University of Texas at Arlington in Partial Fulfillment

of the Requirements

for the Degree of

MASTER OF SCIENCE IN AEROSPACE ENGINEERING

THE UNIVERSITY OF TEXAS AT ARLINGTON

August 2008
Copyright © by Bryan Dale Mixon 2008

All Rights Reserved

ii
ACKNOWLEDGEMENTS

I would first like to thank my supervising Professor, Dr. Kent Lawrence for all

of his guidance throughout the duration of this project. I would also like to thank Dr.

Wen Chan and Dr. Don Wilson for taking the time to serve on my committee. Also, I

would like to thank the Department of Mechanical and Aerospace Engineering, Dr.

Wilson, Dr. Chan, and Dr. Chudoba for the financial support that they have provided to

me over the last couple of years.

I would like to thank the members of the AVD Lab, Gary Coleman, Amit Oza,

and Nauman Mhaskar. Amit and Gary: You guys brought a little bit of OU to UTA. I

would also like to thank JuneHyun Kim and his family: thanks for making me feel like a

part of your family.

I wish to thank my family for all of their support over the last few years. I could

not have done this without you. To my brother, Brad, for having as bad of taste in Sci-

Fi movies as I do and for being a technical sounding board throughout my project. And

to Kristen Roberts for your insight and advice over the years.

August 8, 2008

iii
ABSTRACT

THE DEVELOPMENT OF A FINITE ELEMENT TOOL FOR THE CALCULATION

OF BEAM CROSS SECTION PROPERTIES

Bryan Dale Mixon, MS

The University of Texas at Arlington, 2008

Supervising Professor: Dr. Kent L. Lawrence

Cross section properties are crucial in the analysis and design of beams as they

play a key role in the determination of the stress state encountered. The development of

an FEM tool in MATLAB for the calculation of beam cross section properties will be

discussed. Properties of interest include general cross section properties as well as the

non-trivial warping-dependent properties for both the unrestrained and restrained cross

section conditions for solid and hollow, open and closed sections.

iv
TABLE OF CONTENTS

ACKNOWLEDGEMENTS....................................................................................... iii

ABSTRACT .............................................................................................................. iv

LIST OF ILLUSTRATIONS..................................................................................... ix

LIST OF TABLES..................................................................................................... xii

Chapter

1. INTRODUCTION ......................................................................................... 1

1.1 Introduction ........................................................................................ 1

1.2 Problem Description ........................................................................... 2

1.3 Previous Work .................................................................................... 4

2. BACKGROUND AND THEORY ................................................................ 6

2.1 Warping Independent Cross Section Properties ................................. 6

2.2 Warping Dependent Cross Section Properties.................................... 8

2.2.1. Unrestrained Warping Case..................................................... 8

2.2.1.1. Saint-Venant Torsion Problem.................................... 8

2.2.1.2. Alternative Formulation of Torsion Problem.............. 19

2.2.2. Restrained Warping Case ........................................................ 22

3. FINITE ELEMENT FORMULATION......................................................... 33

3.1 Elliptical Boundary Value Problem Formulation ............................... 34

v
3.1.1. Weak Solution for Poisson and Laplace Equations................. 35

3.1.2. Discretization of Weak Solution: Ritz-Galerkin Method ........ 37

3.2 Element Development......................................................................... 39

3.2.1. Three Node Triangle: Area Coordinates.................................. 39

3.2.2. Six Node Triangle: Isoparametric Coordinates ....................... 42

3.3 FE Formulation of Discretized Problem (Au=b) ................................ 45

3.3.1. Left Hand Side: Stiffness Matrix ............................................. 45

3.3.1.1. Three Node Triangles.................................................. 46

3.3.1.2. Six Node Triangles...................................................... 47

3.3.2. Right Hand Side:


Volume Force and Boundary Conditions................................ 49

3.3.2.1. Three Node Triangles.................................................. 49

3.3.2.2. Six Node Triangles...................................................... 53

3.4 Beam Cross Section Properties........................................................... 54

3.4.1. Warping Independent Cross Section Properties ...................... 54

3.4.1.1. Three Node Triangles.................................................. 54

3.4.1.2. Six Node Triangles...................................................... 56

3.4.2. Warping Dependent Cross Section Properties......................... 58

3.4.2.1. Three Node Triangles.................................................. 59

3.4.2.2. Six Node Triangles...................................................... 61

4. RESULTING CROSS SECTION PROPERTY FEM TOOL ....................... 63

4.1 Overview of FEM Code- Basic Structure........................................... 64

4.2 Required Inputs................................................................................... 66


vi
4.2.1. EasyMesh Output Files............................................................ 66

4.2.2. BCSP Finite Element Tool Input Files .................................... 68

4.2.3. BCSP Input File Creator.......................................................... 70

4.3 Analysis and Output............................................................................ 71

4.3.1. BCSP Finite Element Tool ...................................................... 72

4.3.1.1. Read Inputs and Symmetry Condition ........................ 72

4.3.1.2. Stiffness Matrix Development .................................... 74

4.3.1.3. Volume Force Development and Boundary


Condition Application ................................................. 77

4.3.1.4. Solution of Linear System:


Stress Function/Warping Function .............................. 80

4.3.1.5. Beam Cross Section Properties ................................... 81

4.3.2. BCSP Problem Conditioning:


Stress Function vs. Warping Function .................................... 85

4.3.2.1. Stress Function ............................................................ 86

4.3.2.2. Warping Function........................................................ 86

5. VALIDATION AND APPLICATION OF BCSP FEM TOOL.................... 88

5.1 Circular Cylinder ................................................................................ 88

5.2 Square Section with One Square Cut-Out .......................................... 90

5.3 Square Section with Two Square Cut-Outs ........................................ 92

5.4 C-Beam Application ........................................................................... 94

5.5 I-Beam Application............................................................................. 95

5.6 Circular Cylinder Application: 3-Node vs. 6-Node Triangles............ 97

vii
5.7 Stress Function vs. Warping Function ................................................ 98

6. CONCLUSIONS AND RECOMMENDATIONS ............................................... 101

Appendix

A. BCSP FINITE ELEMENT TOOL: MATLAB SOURCE CODE................. 104

REFERENCES .......................................................................................................... 114

BIOGRAPHICAL INFORMATION......................................................................... 116

viii
LIST OF ILLUSTRATIONS

Figure Page

1.1 Warping of an I-Beam cross section [2].......................................................... 2

1.2 Illustration of common cross section classifications....................................... 3

1.3 Cross section coordinate system [5]................................................................ 4

2.1 Cross section under torsion about the x-axis


with displacements uy and uz. [5] .................................................................... 9

2.2 Moment Equilibrium on a cross section [5] .................................................... 17

2.3 Coordinate translation for restrained warping [5] ........................................... 23

3.1 Representative three node triangle divided


into area coordinate relations .......................................................................... 40

3.2 Representative six node triangle divided


with curved sides............................................................................................. 43

4.1 Basic logic flow for cross section FEM Tool.................................................. 65

4.2 Main driver program BCSP.m and external


functions .......................................................................................................... 72

4.3 Input file read-in.............................................................................................. 73

4.4 Cross section symmetry factor ........................................................................ 74

4.5 3-Node triangle element stiffness matrix


(stima3.m) [11]................................................................................................ 74

4.6 6-Node element stiffness matrix (stima6.m)................................................... 75

4.7 6-Node element stiffness matrix (stima6.m) (continued)................................ 76

ix
4.8 Global stiffness matrix development for 3-Node and
6-Node triangles (BCSP.m)............................................................................. 77

4.9 Global volume force assembly........................................................................ 78

4.10 Neumann boundary condition application ...................................................... 79

4.11 Warping Normal Derivative (warp.m) ............................................................ 79

4.12 Dirichlet boundary conditions......................................................................... 80

4.13 Solving the linear system of equations............................................................ 81

4.14 3-Node triangle warping independent and


dependent cross section properties.................................................................. 82

4.15 3-Node Centroidal calculations and


Parallel axis theorem ....................................................................................... 82

4.16 Gauss Quadrature Weighting Functions [20, 21]............................................ 83

4.17 6-Node triangle warping independent and


dependent cross section properties.................................................................. 83

4.18 6-Node warping independent and dependent


cross section properties (continued)................................................................ 84

4.19 6-Node Centroidal calculations and Parallel


axis theorem .................................................................................................... 84

4.20 Unrestrained and Restrained Warping Cross


Section Property Calculations ......................................................................... 85

5.1 Circular Cylinder: 3-Node vs. 6-Node study .................................................. 89

5.2 Square cross section with square hollow center.............................................. 91

5.3 Square cross section with two hollow


rectangular cut-outs ......................................................................................... 93

5.4 Channel Beam ................................................................................................. 94

5.5 I-Beam............................................................................................................. 96

x
5.6 3-Node straight-sided and 6-Node
curved-sided elements ..................................................................................... 97

5.7 Circular Cylinder............................................................................................. 97

5.8 Rectangular C-S: Stress Function


vs. Warping Function ...................................................................................... 99

5.9 Convergence of Stress Function vs. Warping


Function on Torsional Constant ...................................................................... 100

xi
LIST OF TABLES

Table Page

3.1 Weighting function and shape function


values for each Gauss point [20, 21] ............................................................... 45

5.1 Circular Cylinder Results: ANSYS and BCSP ............................................... 90

5.2 Square cross section with square cut-out results............................................. 92

5.3 Square cross section with two cut-outs results................................................ 93

5.4 Channel Beam Results .................................................................................... 95

5.5 I-Beam Results ................................................................................................ 96

5.6 Circular Cylinder: 3-Node vs. 6-Node results................................................. 98

5.7 Rectangular C-S: Stress Function vs. Warping Function................................ 100

xii
LIST OF SYMBOLS AND ABBREVIATIONS

A Area
Qy First Moment of the Area about the y-axis
Qz First Moment of the Area about the z-axis
yc Coordinate of Centroid in y-direction
zc Coordinate of Centroid in z-direction
Iy Area Moment of Inertia about the y-axis
Iz Area Moment of Inertia about the z-axis
Iyz Product of Inertia
ω Warping Function
J Torsional Constant
ω* Restrained Warping Function
Γ Warping Constant
ux Displacement in the x-direction
uy Displacement in the y-direction
uz Displacement in the z-direction
θ Angle of Twist Per Unit Length
λ Lame’s Constant
G Shear Modulus
γ xy Shear Strain Component on the x-Plane in the y-Direction

γ xz Shear Strain Component on the x-Plane in the z-Direction

γ yz Shear Strain Component on the y-Plane in the z-Direction

τ xy Shear Stress Component on the x-Plane in the y-Direction

τ xz Shear Stress Component on the x-Plane in the z-Direction

xiii
τ yz Shear Stress Component on the y-Plane in the z-Direction

pvx Body Forces per Unit Volume in the x-Direction


pvy Body Forces per Unit Volume in the y-Direction
pvz Body Forces per Unit Volume in the z-Direction
nx Unit Vectors Normal to the x-Plane surface
ny Unit Vectors Normal to the y-Plane surface
nz Unit Vectors Normal to the z-Plane surface
px Surface Forces per Unit Area in the x-Direction
py Surface Forces per Unit Area in the y-Direction
pz Surface Forces per Unit Area in the z-Direction
ε Strain
σ Bending Stress
u Displacement
E Young’s Modulus
∇2 Laplacian Operator
e Change in Unit Volume per Unit Volume or Dilatation

n Normal Unit Vector

u Displacement Vector
Mx Moment About the x-Axis for a Cross Section of Arbitrary Shape
GJ Torsional Stiffness
ψ ( y, z ) Prandtl stress function
Cyz Centroidal Coordinate Axis
Oy*z* Coordinate Axis about an Unknown Point O
x* Restrained Warping Coordinate of x-Axis
y* Restrained Warping Coordinate of y-Axis
*
z Restrained Warping Coordinate of z-Axis
xθ Rotation of the Cross Section a Distance x Down the Beam
u *x Warping Displacement in x-Direction about Shear Center

xiv
u *y Warping Displacement in y-Direction about Shear Center

u *z Warping Displacement in z-Direction about Shear Center


Qω First Moment of Warping
Izω Warping Dependent Cross Section Property Called the Sectorial
Product of Area About the Z-Axis
Tω Warping Torque
σω Warping Normal Stress
Ne Vector of Elemental Shape Functions
n Number of Gauss Points Used
Je Determinant of the Jacobian Matrix

Ke Element Stiffness Matrix


Le Length of the line defined by the node pair of an element that is
touching a boundary
g Warping Normal Derivative
be Element Volume Force
Qy e Elemental First Moment of the Area about the y-Axis

Qz e Elemental First Moment of the Area about the z-Axis

I ye Elemental Moment of Inertia about the y-Axis

I ze Elemental Moment of Inertia about the z-Axis

I yz e Elemental Product of Inertia

Qω e Elemental First Moment of Warping


ys y-Coordinate for the Shear Center
zs z-Coordinate for the Shear Center

xv
CHAPTER 1

INTRODUCTION

1.1 Introduction

Beam cross sections come in many shapes and sizes, being of either open or

closed configuration, having singly or multiply connected regions. Regardless of these

possibilities, there are several fundamental properties associated with the cross section

of a given beam which are necessary for design and analysis.

These fundamental properties appear in their most basic form as general cross

section properties such as area, moments of inertia, and centroidal coordinates. More

complicated, however, are those properties derived when considering problems of

torsion and transverse shear. For the purpose of this thesis, only general cross section

properties and those properties due to torsion are discussed.

Augustin Cauchy first discovered experimentally that cross sections of non-

circular shape experience an out-of-plane displacement called warping (see Figure 1.1),

a contrary result to the previous assumptions made by C.A. Coulomb who suggested

that cross sections remain plane under torsion [1]. Building upon this result, Barrie de

Saint-Venant, a student of Navier, applied Cauchy’s breakthrough to Coulomb's torsion

analysis, describing the axial displacement due to torsion in the form of a function [1].

As a result of the Saint-Venant torsion problem, several important cross sectional

properties appear, such as the torsion and warping constants.

1
Figure 1.1: Warping of an I-Beam cross section [2]

Analytical solutions for these cross section properties exist for simple cross

section shapes but become increasingly difficult for multiply connected cross sections

and cross sections of arbitrary shape. This is especially true for the properties obtained

in the problem of torsion. Specifically, the fact that the warping function is determined

from the solution of a second order partial differential Laplace equation for which

explicit solutions are rare makes the problem complicated [3,4]. Thus, in order to

obtain the fundamental cross section properties for a torsional analysis, it is prudent to

numerically approximate these properties over a given cross section. To this end, such

a task is accomplished proficiently through the use of the finite element method.

1.2 Problem Description

The purpose of this project is to create a finite element tool to provide a

capability for the calculation of fundamental cross section properties. It must handle

open or closed cross sections of arbitrary shape that are singly or multiply connected.

This tool is to be implemented in the MATLAB programming environment to

determine beam cross section properties in the case of pure torsion.

2
Examples of such cross sections are the square section (singly connected, closed

section), the hollow cylinder (multiply connected, closed section), and the I-Beam

(singly connected, open section) as seen in the Figure 1.2.

Figure 1.2: Illustration of common cross section classifications

The cross section properties of interest range of typical cross section properties

to those due to warping on a cross section. The following listing briefly summarizes

these desired properties.

General cross section properties of interest are [5]:

Area (A), first moments of the area (Qy and Qz), centroid coordinates (yc and zc),

area moments of inertia (Iy and Iz), and product of inertia (Iyz).

Warping dependent cross section properties of interest are [5]:

Warping function (ω), torsional constant (J), restrained warping function (ω∗),

and warping constant (Γ)

The following figure defines the coordinate system used for these properties.

3
Figure 1.3: Cross section coordinate system [5]

1.3 Previous Work

Finite element tools capable of the requirements set forth in the previous section

have been developed previously. Perhaps most notably is the powerful finite element

modeling tool known as ANSYS. Within the program is a built-in module called Beam

Tool. The Beam Tool module has 11 common cross sections and the ability to

manually create uncommon cross sections. In both cases, all of the necessary cross

section properties are calculated based on the solution of the warping function [6].

Another example of an FEM tool is that developed by Walter Pilkey, created in the

FORTRAN programming language. First, the user provides an input file describing the

geometric data defining the cross section. After meshing the geometry, the mesh

generator applies nodes and elements to the geometry, and then the FEM code solves

for the desired properties and stresses generated by a given external force and/or

moment [5].

The remaining chapters of this paper detail the theory and development of a

finite element tool to calculate beam cross section properties under torsional loading

4
conditions. Chapter 2 provides a brief background to the theoretical framework. Chapter

3 outlines the finite element formulation of the theory discussed in the previous chapter.

Chapter 4 offers an overview of the finite element code developed in MatLab. Chapter

5 presents validation and application studies, focusing on key cross section types,

followed by their results in order to verify the performance of the tool developed in

Chapter 4. Chapter 6 discusses conclusions and offers recommendations for future study

in this area.

5
CHAPTER 2

BACKGROUND AND THEORY

Before a finite element tool can be created to calculate the cross section

properties for beams, it is first necessary to develop an understanding of the underlying

theory for the problem at hand. As outlined in Chapter 1, there are several cross section

properties of interest. These properties break down into two sections: warping

independent and warping dependent. The warping independent cross section properties

are more common in everyday use such as area, moments of inertia, and the like. The

warping dependent cross section properties break down further into two sub-sections,

unrestrained warping and restrained warping. The unrestrained warping case is

presented as the Saint-Venant torsion problem and represents the foundation for the

warping dependent properties. The restrained warping case will demonstrate the

adjustments necessary to augment the Saint-Venant theory for open cross sections.

2.1 Warping Independent Cross Section Properties

For any cross section, there are several cross section properties that are

independent of warping. These properties are most often defined by the following

equations; these have been obtained from References 5 and 7. See Figure 2.1 below.

The area, A, of a geometric body is defined by the following integral:

A = ∫ dA (2.1)
A

6
where the operator, dA, represents a differential area on the body and has coordinates y

and z.

The first moments of the area, Qy and Qz, are defined as:

Q y = ∫ z dA (2.2)
A

Q z = ∫ y dA (2.3)
A

in which y and z are the coordinates of the differential area dA [5,7]. The first moments

of the area are essential in determining the location of centroid coordinates yc and zc, as

seen in Equations 2.4 and 2.5 below [5,7].

Qz
yc = (2.4)
A

Qy
zc = (2.5)
A

The moments of inertia for a plane area with respect to the y and z axes are

given by the following area integrals: [5,7]

I y = ∫ z 2 dA (2.6)
A

I z = ∫ y 2 dA (2.7)
A

where the differential area is multiplied by the square of its coordinate in z or y

respectively. Due to this, the moments of inertia are sometimes referred to as the

second moments of the area [7].

I y z = ∫ y z dA (2.8)
A

7
2.2 Warping Dependent Cross Section Properties

The warping dependent cross section properties outlined in this section are

classified by either the unrestrained warping case or the restrained warping case.

2.2.1. Unrestrained Warping Case

The problem of torsion is outlined in this subsection. Two approaches will be

discussed. The Saint-Venant warping function approach is presented first, followed by

the Prandtl stress function approach.

2.2.1.1. Saint-Venant Torsion Problem

Initially, the problem of torsion was approached by C.A. Coulomb based on the

assumption that cross sections under torsion will remain plane after deformation, thus

only experiencing a rotation about the axis of twist [1]. In other words, the displacement

perpendicular to the cross section is zero. However this method proved to be accurate

only in beams of circular cross sections [5]. Building upon this idea, Barrie de Saint-

Venant made a slightly different assumption. He proposed that the cross section of a

beam, when acted upon by pure torsion about the axis of twist, will not only rotate

about the axis of twist, as theorized by Coulomb, but will also experience a

displacement in the direction perpendicular to the beam cross section. The out of plane

warping experienced during pure torsion would also, when viewed as a projection back

onto the original plane, retain the original shape of the cross section [5].

8
Figure 2.1: Cross section under torsion about the x-axis with displacements uy and uz. [5]

The Figure 2.1 shows a cross section under torsion about the x-axis, which is

also the axis of twist. For a specific point on the cross section, the displacement in the y

and z directions is approximated by the following equations:

u y ≈ − z xθ (2.9)

u z ≈ y xθ (2.10)

where θ is the angle of twist per unit length and xθ is the rotation of a section of the

cross section at a distance x down the beam [5].

Coulomb’s work shows that beams of circular cross section under pure torsion

will only rotate, while having no displacement axially (warping). For all non-circular

cross sections, experimental results demonstrate that an axial displacement occurs and

that this displacement is relatively constant throughout the length of the beam [5]. This

implies that axial displacement caused by torsion is governed by some function that is

dependent on the y and z directions but independent of the axial (x) direction.

9
Consequently, Saint-Venant proposed a relation describing axial displacement, ux,

across a given cross section as described by Equation 2.11 below.

u x = θ ω ( y, z ) (2.11)

This warping displacement is proportional to the angle of twist per unit length, θ, and a

function describing the out of plane warping of the cross section, ω(y,z). The unknown

function is referred to as the warping function [5].

As expected, the main task of the problem of torsion is the mathematical

formulation and eventual solution of the unknown warping function, ω(y,z). Saint-

Venant took a displacement formulation approach to this solution; its derivation is

shown here, as taken from Reference 5.

1. Governing Equations of Linear Elasticity

The basis of Saint-Venant’s approach from the case of linear elasticity for the

problem of torsion begins with the strain-displacement and stress-strain relations,

equations of equilibrium, and the surface force equations on a body.

From Reference [5]:

Strain Displacement Relations:

∂u x ∂u y ∂u
εx = , εy = , εz = z (2.12)
∂x ∂y ∂z

∂u y ∂u x ∂u ∂u ∂u ∂u y
γ xy = + , γ xz = z + x , γ yz = z + (2.13)
∂x ∂y ∂x ∂z ∂y ∂z

Stress-Strain Relations:

σ x = λe + 2Gε x , σ y = λe + 2Gε y , σ z = λe + 2Gε z (2.14)

10
τ x y = Gγ x y , τ y z = Gγ yz , τ x z = Gγ x z (2.15)

where λ is Lame’s constant, G is the shear modulus, and the dilation

∆v
e = εx +εy +εz = .
v

The equations of equilibrium which govern the relationship between internal

forces and several differential equations of the stresses are:

∂σ x ∂τ x y ∂τ x z
+ + + pv x = 0 (2.16)
∂x ∂y ∂z

∂τ x y ∂σ y ∂τ y z
+ + + pv y = 0 (2.17)
∂x ∂y ∂z

∂τ x z ∂τ y z ∂σ z
+ + + pv z = 0 (2.18)
∂x ∂y ∂z

where pvx, pvy, and pvz are the body forces per unit volume.

The equations governing the surface forces on a body are given by:

p x = σ x nx + τ x y n y + τ x z nz (2.19)

p y = τ x y nx + σ y n y + τ y z nz (2.20)

p z = τ x z nx + τ y z n y + σ z nz (2.21)

where nx, ny, and nz are the unit vectors normal to the surface and px, py, and pz are the

surface forces per unit area [5].

To this point, the strain-displacement relations for linear elasticity have been

shown in terms of the displacements. Next, the strains are related to the stresses through

11
the stress-strain relations. Now it is possible, using these relations, to represent the

stresses in terms of the displacements ux, uy, and uz. In other words:

ε = Du , σ = E ε , σ = E Du (2.22)

where ε, σ, and u are vectors containing the strains, stresses, and displacements

respectively, E is Young’s Modulus, and D is a matrix containing all of the differential

operators. Introducing Equation 2.22 into the equations of equilibrium provides a set of

partial differential equations in terms of the displacements. When shown in vector

notation, the equations of equilibrium appear as:

D T σ + pv = 0 (2.23)

Substituting the results of Equation 2.22 into the equilibrium equations above will

result in the following modified equations of equilibrium:

(λ + G ) ∂e + G ∇ 2 u x + pv x =0 (2.24)
∂x

(λ + G ) ∂e + G ∇ 2 u y + pv y =0 (2.25)
∂y

(λ + G ) ∂e + G ∇ 2 u z + pv z =0 (2.26)
∂z

∂2 ∂2 ∂2 ∆v
with ∇ = 2 + 2 + 2 , e = ε x + ε y + ε z =
2

∂x ∂y ∂z v

where ∇ 2 is the Laplacian operator and e is the change in unit volume per unit volume,

or dilatation.

2. Application to Pure Torsion

12
The next step is to apply the background outlined in the previous section to the

problem of torsion based on the displacement conditions assumed by Saint-Venant in

Equations 2.9 thru 2.11. Beginning again with the strain-displacement relations

(Equations 2.12 and 2.13) and applying the assumed displacement conditions, the

following is obtained [5]:

∂u x
εx = = θ 'ω ( y , z ) = 0 (2.27)
∂x

∂u y
εy = =0 (2.28)
∂y

∂u z
εz = =0 (2.29)
∂z

∂u y ∂u x  ∂ω 
γ xy = + = θ  − z  (2.30)
∂x ∂y  ∂y 

∂u z ∂u x  ∂ω 
γ xz = + = θ  + y  (2.31)
∂x ∂z  ∂z 

∂u z ∂u y
γ yz = + =0 (2.32)
∂y ∂z

where the strain-displacement equation for εx is equal to zero since θ is constant.

With σx, σy, σz, and τyz equal to zero, the stress-strain relations for the case of

pure torsion (Equations 2.14 and 2.15) are greatly simplified, leaving only shear stress

on the x-face in the y and z directions.

τ x y = Gγ x y , τ x z = Gγ x z (2.33)

13
Substituting the shear strain equations from Equation 2.30 and 2.31 above provides the

following stress-strain equations in terms of the warping function:

 ∂ω 
τ x y = Gγ x y = Gθ  − z  (2.34)
 ∂y 

 ∂ω 
τ x z = Gγ x z = Gθ  + y  (2.35)
 ∂z 

Finally, returning to the equations of equilibrium (Equations 2.24 thru 2.26), the

effects of the displacement conditions with the unknown warping function must be

considered. Note that for the pure torsion problem, all body forces will be zero [5].

Equations 2.36 and 2.37 below restate the equations of equilibrium for the y and z

components.

(λ + G ) ∂e + G ∇ 2 u y + pv y =0 (2.36)
∂y

(λ + G ) ∂e + G ∇ 2 u z + pv z =0 (2.37)
∂z

Realizing that,

∂u x ∂u y ∂u z
e = εx +εy +εz = + + =0 (2.38)
∂x ∂y ∂z

and that the Laplacian of uy and uz are likewise equal to zero, Equations 2.36 and 2.37

will be equally satisfied, vanishing. Examining the x component of the equilibrium

equations for the pure torsion problem with the displacement condition in the axial

direction taken as,

14
u x = θ ω ( y, z ) (2.39)

Equation 2.24 becomes

∂e x
(λ + G ) + G∇ 2 (θ ω ( y, z )) = 0
∂x

∇ 2ω ( y, z ) = 0

∂ 2ω ∂ 2ω
∇ 2ω ( y , z ) = + =0 (2.40)
∂y 2 ∂z 2

With the formulation of Equation 2.40, the unknown warping function for the

problem of torsion is now defined as the solution to this partial differential equation [1,

5, 8]. Specifically, a PDE of this nature is described as Laplace’s equation, the solution

of which is a harmonic function [5].

Having identified the Laplace equation above, it is necessary to inspect the

surface conditions defined in Equations 2.19 thru 2.21, with respect to the problem at

hand. These surface/boundary condition equations must be rewritten in terms of

displacements to completely define the displacement formulation. This is accomplished

by first writing the surface equations in terms of strains using the material laws, and

then using the strain-displacement relations to express the strains in terms of the

displacements. The result of which, as shown in Reference [5], is:


− −∂u
p x = λ e n x + G n ⋅ ∇u x + G n ⋅ (2.41)
∂x


− −∂u
p y = λ e n y + G n ⋅ ∇u y + G n ⋅ (2.42)
∂y

15

− − ∂u
p z = λ e n z + G n ⋅ ∇u z + G n ⋅ (2.43)
∂z

− −
where e is again equal to zero, n is the normal unit vector, and u is the displacement

vector.


For the condition that the x component of the unit normal vector, n , is equal to

zero, and with no applied forces on the body in question (i.e. px, py, and pz = 0),

Equation 2.42 and Equation 2.43 vanish. The x-direction of the surface condition, under

the same circumstances, reduces to the form

∂u y ∂u x ∂u y ∂u
ny + nz + ny + nz z = 0 (2.44)
∂y ∂z ∂x ∂x

Collecting terms and comparing to the strain-displacement equations, the following

boundary condition is obtained:

 ∂ω   ∂ω 
n y  − z  + nz  + y = 0
 ∂y   ∂z 

∂ω ∂ω
ny + n z = zn y − yn z
∂y ∂z

∂ω ∂ω ∂ω
where ny + nz =
∂y ∂z ∂n

∂ω
= zn y − yn z (2.45)
∂n

The boundary condition of Equation 2.45 is of the Neumann type since it is in the form

of a normal derivative.

16
Having now devised the boundary condition requirement necessary for the

Laplace boundary value problem (BVP) for the warping function for pure torsion, it is

now prudent to inspect the effect of torsion on the body. The unit normal vectors found

at either end of a given beam are n x = ±1 and n y = n z = 0 . Application to the surface

condition equations yields:

px = 0 (2.46)

py =τ xy (2.47)

pz = τ xz (2.48)

which results in a torque on the body.

Moment equilibrium requirements for the cross section depicted in Figure 2.2

provide the equation below for the moment, Mx.

M x = ∫∫ (τ x z y − τ x y z )dy dz (2.49)
A

Figure 2.2: Moment Equilibrium on a cross section [5]

Substituting Equations 2.34 and 2.35 into the above gives:

17
 ∂ω   ∂ω  
M x = Gθ ∫  + y  y −  − z  z  dA (2.50)
A 
∂z   ∂y  

which defines the moment about the x-axis for a cross section of arbitrary shape.

The moment equation about the axis of twist for a circular cross section, which

has zero warping, is:

M x = GJθ (2.51)

where J is the polar moment of inertia, (Iy + Iz) [1, 5, 8]. As expected, when zero

warping is assumed in Equation 2.50, the equation for a circular cross section, Equation

2.51, is obtained. Based off this discovery, the integral portion of Equation 2.50 is now

identified as one of the desired cross section properties known as the torsional constant,

J.

 ∂ω   ∂ω  
J = ∫  + y  y −  − z  z  dA (2.52)
A 
∂z   ∂y  

The product GJ is labeled torsional stiffness and is a measure of a cross section’s

resistance to torsion.

The Saint-Venant torsion problem is now fully formulated. Solving the partial

differential Laplace equation (Equation 2.40) with the boundary condition given by

Equation 2.45, the warping function is obtained. With the warping function solved, the

torsional constant, J, of Equation 2.52 can be found. Section 2.2.2 will later supplement

this formulation with the introduction of another cross section property: the warping

constant Γ.

18
It is important to note that the warping function approach, outlined by the Saint-

Venant torsion problem, is only valid for certain types of cross sections, under certain

torsional loading scenarios. Saint-Venant made the assumption that cross sections are

free to warp without restraint (i.e. unrestrained warping). If, for example, the beam has

external supports, is not prismatic, or has a distributed torsional load, the previously

made assumption is not satisfied [5]. Even so, it is seen that for closed, solid or hollow

cross sections, the warping function approach yields acceptable, approximate results

[5]. Unfortunately, for open thin-walled cross sections, it is necessary to take some

additional steps. Section 2.2.2 explains these steps and the additional properties that

must be determined to characterize the cross section.

2.2.1.2. Alternative Formulation of Torsion Problem

In the previous subsection, Saint-Venant’s approach to the torsion problem was

formulated. The resulting equations formed a boundary value problem of the Laplace

equation in terms of the warping function ω(y,z) with a Neumann type boundary

condition. According to reference [4], “the nature and form of the boundary

conditions” relating to this solution are inconvenient in analysis for cross sections of

irregular shape. From the theory of analytic functions, an alternative approach was

developed by Prandtl, named the Prandtl Stress Function,ψ ( y, z ) , referred to as the

force formulation. The main advantage given here is the existence of greatly simplified

boundary conditions [1].

The stress function,ψ ( y, z ) , may be defined as

19
∂ψ ∂ψ
τ xy = , τ xz = − (2.53)
∂z ∂y

with σx, σy, σz, τyz, and body forces equal to zero for the case of pure torsion about the

axis of twist, x-axis. The equations of equilibrium reduce to the following:

∂τ x y ∂τ x z
+ =0 (2.54)
∂y ∂z

∂τ y x ∂τ zx
= =0 (2.55)
∂x ∂x

The goal now is to determine the value of the Prandtl stress function over the cross

section. From the compatibility equations for linear elasticity and the shear strain

relations for the problem of torsion, Prandtl obtained the following two relations for the

stress function [5]

∂ 2
∇ψ =0 (2.56)
∂y

∂ 2
∇ψ =0 (2.57)
∂z

∂2 ∂2 ∂2
with ∇ 2 = + +
∂x 2 ∂y 2 ∂z 2

which, by examination, leads to the conclusion that ∇ 2ψ must be equal to a constant.

Relating the above equations to the original stress function relation for the shear stresses

it can be seen that

∂ 2ψ ∂τ  ∂ 2ω 
= − = − Gθ  + 1 (2.58)
∂y 2
∂y  ∂ y∂ z 

20
∂ 2ψ ∂τ  ∂ 2ω 
= = Gθ  − 1 (2.59)
∂z 2
∂y  ∂z ∂ y 

Summing these two results provides the following partial differential equation.

∇ 2ψ = −2Gθ (2.60)

A PDE of this type is called Poisson’s equation.

Boundary conditions for the stress function formulation are derived from the

same equation as for the warping function approach, Equation 2.45, resulting in the next

equation.

∂ψ d y ∂ψ d z ∂ψ
+ = =0 (2.61)
∂ y d s ∂z d s ∂ s

From this equation it follows that ψ must be a constant along the boundary of the cross

section. For simplicity the value of the stress function is taken to be zero. In a similar

manner to the Saint-Venant torsion problem in the previous subsection, the equation

defining the moment about the axis of twist in terms of the stress function is

 ∂ψ ∂ψ 
M x = ∫∫  − y− z  dy dz (2.62)
A 
∂y ∂z 

After integration this becomes

M x = 2∫∫ψ dy dz (2.63)
A

Finally, the torsional constant may be defined for Prandtl’s Stress function approach as

Mx 2
Gθ Gθ ∫A
J= = ψ dA (2.64)

21
2.2.2. Restrained Warping Case

The theory outlined in Section 2.2.1.1 regarding Saint-Venant’s Theory of

torsion is based largely on the assumption that cross sections are left free to warp, or in

other words, encounter unrestrained warping [5]. This free warping assumption

provides valid results for solid and closed thin-walled cross sections, but poses a

problem for open thin-walled cross sections [5, 16]. When open thin-walled cross

sections experience a torsional load, the deflections encountered are much higher in

comparison, due to a much smaller torsional rigidity, GJ. As a result, the effects of

axial stress, σx, cannot be ignored as was done previously in Saint-Venant’s theory [5].

In Saint-Venant’s unrestrained warping approach, the location of the axis of

twist is not important, since it has no direct affect on torsional shear stress calculations

[5]. In restrained warping, however, the location of the axis of twist is not arbitrary and

must be determined. The process used to formulate restrained warping proceeds much

the same as Saint-Venant’s unrestrained warping, but with the inclusion of the unknown

location of center of twist. The process, as described in reference [5], follows.

Firstly, let Figure 2.3 define a centroidal coordinate axis Cyz and also a

coordinate axis Oy*z* about an unknown point O.

22
Figure 2.3: Coordinate translation for restrained warping [5]

Relating y* to y and z* to z in Figure 2.3 provides the following equations

defining the coordinates y* and z*.

y* = y − b (2.65)

z* = z − c (2.66)

The goal of unrestrained warping is to arrive at a value for the warping function,

ω(y,z), about the defined coordinate axis. In the case of restrained warping, let us find

the warping function about the Oy*z* coordinate axis resulting in the following notation.

ω * ( y* , z* ) (2.67)

Equations defining displacements for an arbitrary point on the cross section

must be expressed for the restrained warping case. Using the coordinate equations for

y* and z* and Equations 2.9-2.11, the displacement relations for restrained warping

appear as

u *y = − z * xθ = − xθ ( z − c) (2.68)

23
u *z = y * x θ = x θ ( y − b) (2.69)

u x* = θ ω * ( y * , z * ) (2.70)

where θ is the angle of twist per unit length and xθ is the rotation of the cross section a

distance x down the beam [5].

Based on these displacement conditions and the fact that for pure torsion, in

restrained warping, σy, σz, and τyz are equivalently zero (σx is non-zero, unlike in the

previous sections, and will be addressed soon), the stress state for the cross section can

be expressed. This is accomplished through the manipulation of the stress-strain

relations and the strain-displacement equations resulting in

 ∂ω * 
τ x y = Gγ x y = Gθ  − z *  (2.71)
 ∂y 

 ∂ω * 
τ x z = Gγ x z = Gθ  + y *  (2.72)
 ∂z 

Through examination of the equilibrium equations (Equations 2.24 thru 2.26) for

the given displacement conditions set in this subsection, it is noticed that both the y* and

z* components are satisfied, leaving only the x* component. Like in the unrestrained

warping approach, the result is found to be Laplace’s equation.

∇ 2ω * = 0 (2.73)

This time, however, the equation is referenced to the unknown location of the

coordinate axis Oy*z*.

Equation 2.45 of Section 2.2.1.1 characterizes the boundary condition necessary

for a cross section with zero body forces acting on its surface. Taking into account the

24
displacement conditions y* and z*, found in Equation 2.68 and Equation 2.69, the

boundary conditions for restrained warping are obtained.

 ∂ω *   ∂ω * 
n y  − z *  + n z  + y *  = 0 (2.74)
 ∂y   ∂z 

Inserting the equations for y* and z*, the boundary condition is expanded as follows

 ∂ω *   ∂ω * 
n y  + c  + n z  − b  = zn y − yn z (2.75)
 ∂y   ∂z 

Equation 2.75 may be conveniently simplified and rewritten as the following normal

derivative

∂ *
∂n
( )
ω + cy − bz = zn y − yn z (2.76)

Reference [5], suggests that since ω* satisfies the Laplace equation of Equation

( ) ( )
2.73, so too must ω * + cy − bz . In addition, ω * + cy − bz satisfies the boundary

condition of Equation 2.76, which, in comparison, shares the same form as the

boundary condition relation for unrestrained warping (Equation 2.45). Implied from

this observation is that ω* and ω at most only differ by a constant [5]. In equation form

this translates to:

ω * ( y, z ) + cy − bz = ω ( y, z ) + C (2.77)

and isolating ω* yields

ω * ( y, z ) = ω ( y, z ) − cy + bz + C (2.78)

Next, it is necessary to further augment the warping displacement relation of

Equation 2.70 to be not only a function of y and z, but x as well.

25
u x* = θ ( x) ω * ( y , z ) (2.79)

When compared to the warping displacement assumed for unrestrained warping, the

difference here is that the angle of twist per unit length, θ, now varies as a function of

axial location and the restrained warping function, ω*, is incorporated. The restrained

warping function, however, is defined with respect to an undetermined reference point

O on the cross section which marks the axis of twist.

It is important to remember that in the case of restrained warping, axial stresses

develop that can no longer be ignored as they once were in unrestrained warping. For

this reason, Equation 2.79 includes the rate of angle of twist that varies as a function of

x. This inclusion allows for the development of axial stress or warping normal stress, in

the restrained warping approach as follows:

σ x = σ ω = Eε x (2.80)

with

∂u x ∂θ ( x)
εx = = (ω * ) (2.81)
∂x ∂x

Inserting εx from Equation 2.81 into Equation 2.80 yields the warping normal stress,

Equation 2.82.

∂θ ( x)
σ x = E (ω * ) (2.82)
∂x

Since the beam is in an assumed state of pure torsion, the only normal stresses

encountered are those generated by the warping of the cross section. Therefore, the

26
normal stresses that develop must be equivalent to a zero force-couple system where the

following conditions must be satisfied [5].

∫σ
A
x dA = 0 M z = − ∫ yσ x dA = 0
A
M y = ∫ zσ x dA = 0
A
(2.83)

Substitution of Equation 2.82 for σx into each of the above integrals will provide the

following set of equations in terms of the restrained warping function, ω*.

∫ ω dA = 0
*
(2.84)
A

M z = ∫ yω * dA = 0 (2.85)
A

M y = ∫ zω * dA = 0 (2.86)
A

Beginning with Equation 2.84, expanding ω* by Equation 2.78 yields the

following:

∫ ω dA = ∫ ω dA − c ∫ y dA + b ∫ z dA + CA = 0
*
(2.87)
A A A A

The first integral term on the right hand side of the equation is defined as the warping

dependent cross section property called the first moment of warping, Qω, and presented

here as Equation 2.88 [5]. The second and third integral terms on the right hand side

are seen to be zero.

Qω = ∫ (ω dA) (2.88)
A

Equation 2.87 may now be written as the combination of this geometric property and

the constant, C, times the cross section area, A. Thus,

27
∫ ω dA = Qω + CA = 0
*
(2.89)
A

which implies that


C=− (2.90)
A

With the relation for the constant C above, Equation 2.78 now becomes


ω * ( y, z ) = ω ( y, z ) − cy + bz − (2.91)
A

leaving the unknown constants, b and c, to be determined.

Moving now to address the moment condition, Mz, of Equation 2.85 by

substituting the complete relation for ω* yields the following:

M z = ∫ yω *dA = ∫ ( y ω )dA − c ∫ ( y 2 )dA + b ∫ ( y z )dA = 0 (2.92)


A A A A

where the second and third integral terms represent the familiar moment of inertia, Iy,

and product of inertia, Iyz. The first integral in Equation 2.92 is defined as the warping

dependent cross section property called the sectorial product of area, Iyω, about the y-

axis, which is given as [5]:

I yω = ∫ y ω dA (2.93)
A

The moment condition, My, of Equation 2.86, when submitted to the same

treatment as Equation 2.85 provides the definition of another warping dependent cross

section property. It follows that:

M y = ∫ z ω *dA = ∫ ( z ω )dA − c ∫ ( y z )dA + b ∫ ( z 2 )dA = 0 (2.94)


A A A A

28
where the second and third integrals of this relation represent the familiar moment of

inertia, Iz, and product of inertia, Iyz. The first integral of Equation 2.94 is defined as the

warping dependent cross section property called the sectorial product of area, Izω, about

the z-axis, which is given as [5]:

I zω = ∫ z ω dA (2.95)
A

Inserting the defined warping dependent cross section properties (Q∗, Iy∗, Iz∗)

into Equations 2.92 and 2.94 accordingly, yields relations for the unknown constants b

and c which may now be determined.

M z = I yω − cI z + bI yz = 0 (2.96)

M y = I zω − cI yz + bI y = 0 (2.97)

Solving for b and c provides the following expressions:

I yz I yω − I z I zω
b = ys = (2.98)
I y I z − I yz2

I y I yω − I yz I zω
c = zs = (2.99)
I y I z − I yz2

which are equivalent to the Trefftz shear center approximation equations [5].

Incorporating the shear center constants back into the restrained warping

equation, Equation 2.91, the final development of the restrained warping function

equation is reached.


ω * ( y * , z * ) = ω ( y, z ) − z s y + y s z − (2.100)
A

29
In Section 2.2.1.1, it was seen that the Saint-Venant torque, developed in terms

of the warping function, contained what was defined as the torsional constant, J.

Referring back to Equations 2.50 and 2.52 this appeared as:

 ∂ω   ∂ω  
M x = Gθ ∫  + y  y −  − z  z  dA (2.101)
A 
∂z   ∂y  

 ∂ω   ∂ω  
J = ∫  + y  y −  − z  z  dA (2.102)
A 
∂z   ∂y  

In much the same fashion, another important cross section property is defined

from the investigation of shear forces on the cross section. Specifically, when a shear

force acts at any given point on the cross section it is appropriate to replace it with a

force-couple about the shear center which is equivalent to the original shearing force.

The moment produced by the force-couple is referred to as the warping torque, Tω, and

is added to any pure torque that might also be applied to the cross section. The

complete derivation of the warping torque may be found in reference [5] and is defined

here in its final form as:

Tω = − Eθ '' ∫ (ω * ) 2 dA (2.103)
A

where E is Young’s modulus, θ ' ' is the second derivative of the angle of twist per unit

length. The integral of the square of the restrained warping function is identified as

another desired cross section property known as the warping constant, Γ [5].

Γ = ∫ (ω * ) 2 dA (2.104)
A

30
The warping constant Γ of Equation 2.104 is crucial in calculations for restrained

warping. It is useful when calculating the warping torque, Tω, the warping normal

stress, σω, and other important parameters of the restrained warping analysis [5].

Inserting the restrained warping function, ω∗, from Equation 2.100 into Equation 2.104

one obtains:

2
 Q 
Γ = ∫  ω ( y, z ) − z s y + y s z − ω  dA (2.105)
A
A 

Once expanded, Equation 2.105 becomes:

 2  Qω2
Γ =  ∫ ω dA  − − 2 z s I yω + 2 y s I zω + z s2 I z − 2 y s z s I yz + y s2 I y (2.106)
A  A

where the integrated term represents the final warping dependent cross section property

of interest, Iω.

I ω = ∫ ω 2 dA (2.107)
A

Adding this newly defined property to Equation 2.106 and applying some simplification

leads to the following definition of the warping constant.

Qω2
Γ = Iω − − y s2 I y + 2 y s z s I yz − z s2 I z (2.108)
A

In summary, the restrained warping approach is primarily necessary when

dealing with cross sections of an open nature. In such instances, a component of axial

stress, called the warping normal stress, is present. Thus an alternative formulation of

the warping function is created upon which six important warping dependent cross

section properties were defined. For convenience, these results are listed again below:

31

ω * ( y * , z * ) = ω ( y, z ) − z s y + y s z − (2.109)
A

Qω2
Γ = Iω − − y s2 I y + 2 y s z s I yz − z s2 I z (2.110)
A

Qω = ∫ (ω dA) (2.111)
A

I yω = ∫ y ω dA I zω = ∫ z ω dA I ω = ∫ ω 2 dA (2.112)
A A A

32
CHAPTER 3

FINITE ELEMENT FORMULATION

Chapter 2 culminated in the development of partial differential equations for the

warping function and the stress function approach to the problem of torsion. The

warping function approach is of the form of Laplace’s equation and the stress function

approach is of the form of Poisson’s equation. These partial differential equations are

classified as elliptical boundary value problems, and in both cases an explicit solution is

often difficult or nearly impossible to obtain for many cross sections [3]. With this in

mind it is necessary to formulate an alternate approach to the given problem.

This approach is the finite element method. The finite element method is one of

taking a typically analytical problem and discretizing over the domain into much

smaller subspaces that can be individually solved by numerically approximating the

problem over a set of basis functions, or shape functions, which define the nodes of a

given element on the subspace. In this vein, the following chapter presents the

theoretical foundation upon which a finite element approach may be applied to the

problem of torsion [9].

First, a general weak solution to the Laplace and Poisson equation’s is

presented, followed by the Ritz-Galerkin discretization. Next, the elements used to

cover the domain are developed. Finally, the finite element representation for the

problem of torsion and each of the cross section properties are given.

33
3.1 Elliptical Boundary Value Problem Formulation

The first step toward producing a finite element formulation to the problem of

torsion is to define a general model of the problem. This translates into a general model

for both Laplace’s equation, whose solution provides the warping function, and

Poisson’s equation, whose solution yields the stress function. These two may be

summed up in the following general equations which provide a description of the

problem of torsion [10, 11]:

− ∇ 2 u = f in Ω (3.1)

u = u D on ΓD (3.2)

∂u
= g on ΓN (3.3)
∂w

where Ω denotes the entire domain of interest, while ΓD and ΓN denote the Dirichlet

and Neumann boundaries.

If f is taken to be zero, the above equations define Laplace’s equation, where u is

the warping function of Chapter 2. Equations 3.2 and 3.3 are Dirichlet and Neumann

boundary conditions, respectively. If f equals 2 and the Neumann boundary condition

of Equation 3.3 is assumed to vanish, the set of equations would then represent

Poisson’s equation, where u is the stress function.

With this general problem definition set forth, a variational or weak formulation

of the problem of torsion is constructed in Section 3.1.1. The weak formulation is so

labeled due to the lessened smoothness requirement that is necessary within the

classical solution to boundary value problems; as a result it is easier to have the

34
existence of a solution [10]. In addition, weak formulations for elliptical boundary

value problems, such as the problem of torsion, naturally lend themselves to the

development of Galerkin style numerical methods, which will be presented in Section

3.1.2 [10].

3.1.1. Weak Solution for Poisson and Laplace Equations

The general model for the prescribed problem of torsion is defined by Equations

3.1 thru 3.3. In order to produce a weak formulation for this case, begin by making the

assumption that the Dirichlet boundary condition is zero on the boundary.

− ∇ 2 u = f in Ω (3.4)

u = 0 on ΓD (3.5)

∂u
= g on ΓN (3.6)
∂w

Next, multiply Equation 3.4 by an arbitrary, smooth test function, w, and

integrate over the entire domain using Green’s Theorem resulting in the following

equation [10]:

∫ (∇u ⋅ ∇w ) dΩ = ∫ f w dΩ + ∫ g w dΓ
Ω Ω ΓN
N (3.7)

For this new relation to make sense, it is necessary that u ∈ H 1 (Ω) and w ∈ H 10 (Ω) ,

where H represents a vector space [10]. Also, remember that the dirichlet condition of

u = u D on ΓD is desired, which is defined to be an element of the following vector

space; u D ∈ H 1 / 2 (ΓD ) [10]. The problem now stands as follows:

u ∈ H 1 (Ω) , u = u D on ΓD

35
∫ (∇u ⋅ ∇w ) dΩ = ∫ f w dΩ + ∫ g w dΓ
Ω Ω ΓN
N (3.8)

for all w ∈ H 1 0 (Ω)

It is important to note that the trial function, u, and the test function, w, do not

lie in the same space. In order to ensure the existence of a unique solution to the

desired problem, as provided by the Lax-Milgram lemma, it is necessary to propose an

alternative formulation. Reference [10] proposes the following method for augmenting

the problem space. Since u D ∈ H 1 / 2 (ΓD ) , and given the relation γ ( H 1 (Ω) = H 1 / 2 (ΓD ) ,

then there exists a function U D such that γ U D = u D . Thus setting u equal to the

following relation

u = v +UD (3.9)

and substituting this back into Equation 3.8, the problem now becomes one of finding

the solution to v as defined below.

v ∈ H 10 (Ω)

∫ (∇v ⋅ ∇w ) dΩ = ∫ f w dΩ + ∫ g w dΓ − ∫ (∇U
Ω Ω ΓN
N

D ⋅ ∇w) dΩ (3.10)

for all w ∈ H 1 0 (Ω)

Since both v and w lie in the same space (i.e. V = H 01 (Ω) ), the Lax-Milgram lemma

proves that a unique solution exists to the problem such that [10, 12]:

v ∈ V , a (v, w) = b ( w) , for all w ∈ V (3.11)

where from Equation 3.10:

36
a (v, w) = ∫ (∇v ⋅ ∇w ) dΩ (3.12)

and b ( w) = ∫ f w dΩ + ∫ g w dΓN − ∫ (∇U D ⋅ ∇w) dΩ (3.13)


Ω ΓN Ω

represents the weak solution for the general problem of torsion. The next step is to

apply a Galerkin Discretization to Equation 3.11 to fully compose the finite element

formulation of the second order, elliptical, partial differential Laplace and Poisson

equations.

3.1.2. Discretization of Weak Solution: Ritz-Galerkin Method

Instead of minimizing the test function in the weak solution described above

over the vector space V = H 01 (Ω) , it is useful to define a finite dimensional subspace

upon which to minimize the problem. Choose the subspace S D ∈ V , which represents a

smaller or discretized space [13]. The problem now becomes one of seeking

vD ∈ S D

∫ (∇v

D ⋅ ∇wD ) dΩ = ∫ f wD dΩ + ∫ g wD dΓN − ∫ (∇U D ⋅ ∇wD ) dΩ
Ω ΓN Ω
(3.14)

for all wD ∈ S D

Impose {N1 , N 2 ,..., N n } to be a set of basis functions describing the discretized

region S D in the form of [5, 11, 13]:

T
wD = N j w e = w e N Tj (3.15)

As a result, the formulation becomes:

T T
seek v D ∈ S D a (v D , w e N Tj ) = b ( w e N Tj ) j =1,2, ..., n (3.16)

37
or

vD ∈ S D


∫ (∇v D ⋅ ∇N Tj ) dΩ = ∫ f N Tj dΩ + ∫ g N Tj dΓN − ∫ (∇U D ⋅ ∇N Tj ) dΩ
Ω ΓN Ω
(3.17)

Next, from the Ritz-Galerkin method, v D and U D are assumed to take on the

form of [13]:

n
vD = ∑ xk N k (3.18)
k =1

n
U D = ∑U k N k (3.19)
k =1

Substitution of Equations 3.18 and 3.19 into Equation 3.17 yields the following final

form of the discretized problem.

n
[ ∫ (∇N k ⋅ ∇N Tj ) dΩ] xk = ∫ f N Tj dΩ + ∫ g N Tj dΓN − ∑U k ∫ (∇N k ⋅ ∇N Tj ) dΩ (3.20)
Ω Ω ΓN k =1 Ω

for j , k =1,2, ..., n

The system of equations above can be expressed in matrix form as

A( N k , N Tj ) xk = b ( N Tj ) (3.21)

where

A = ∫ (∇N k ⋅ ∇N Tj ) dΩ (3.22)

and

n
b = ∫ f N Tj dΩ + ∫ g N Tj dΓN − ∑U k ∫ (∇N k ⋅ ∇N Tj ) dΩ (3.23)
Ω ΓN k =1 Ω

38
Section 3.3 will discuss what Equations 3.22 and 3.23 represent for the problem

of torsion. Before this is done, it is necessary to develop the basis functions used to

describe the subspace of this section. These basis or shape functions will describe the

behavior of small triangles and the nodes contained therein for the entire domain.

3.2 Element Development

Having completed the discretization of the general problem of torsion, important

basis functions have been introduced. These functions, called shape functions, given

as {N1 , N 2 ,..., N n } over the discretized domain, individually define small elements, but

together represent a mesh covering the region [13]. Each element is composed of three

or more node points which are connected to one another by straight or curved lines.

Each node’s displacement is governed by the shape function of the element it resides in

[14].

For the development of a finite element tool to perform the goals set forth in this

project it is necessary to define the elements that will be used. Two different element

types are developed. The first is a straight sided three node triangle and the second is a

six node triangle able to accommodate curved sides.

3.2.1. Three Node Triangle – Area Coordinates

The three node triangle, also called the linear triangle, is the first element to be

developed here. It is composed of three nodes connected by three straight edges [14].

This element is conveniently described by way of area coordinates.

Figure 3.1 illustrates a three node triangle, one node at each vertice, numbered

in a counter-clockwise manner starting from the lower left node, with a single interior

39
point, P(y,z). Point P is used to divide this triangle into three sub-triangles. Area one is

located opposite of node one. Areas two and three are located opposite of their

respective nodes.

Figure 3.1: Representative three node triangle divided into area coordinate relations

Three functions are described using the ratio of each sub-triangle to the total area of the

triangle. These functions are the shape functions which define each node of the linear

three node triangle at each of its nodes [14, 15].

A1 A A
η1 = , η 2 = 2 , η3 = 3 (3.24)
Ae Ae Ae

The areas A1, A2, and A3 above are easily obtained as half the determinant of the

matrix containing the coordinate locations of the vertices.

1 y z
A1 = det 1 y 2 z 2 
1
(3.25)
2
1 y 3 z 3 

1 y z
A2 = det 1 y1 z1 
1
(3.26)
2
1 y 3 z 3 

40
1 y z
1 
A3 = det 1 y1 z1  (3.27)
2
1 y 2 z 2 

The shape function relations of Equation 3.24 may now be rewritten in terms of

the area equations above after, the determinant has been taken and terms have been

collected, resulting in:

A1 1 A2 1
η1 = = [a1 + b1 y + c1 z ] η2 = = [a2 + b2 y + c2 z ]
Ae 2 Ae Ae 2 Ae

A3 1
η3 = = [a3 + b3 y + c3 z ] (3.28)
Ae 2 Ae

where the variables a, b, and c are given by the following relationships in Equation

3.29.

ai = y j z k − y k z j bi = z j − z k ci = − ( y j − y k ) (3.29)

In the above equation i, j, and k represent a specific node of the element and should be

applied as 1-2-3 consecutively. In practice this means that if a1 is sought, then j=2 and

k=3; if a2 is sought, then j=3 and k=1.

The area for the element may be represented similarly to Equations 3.25-3.27

except the matrix should contain all three of the nodal coordinates as:

1 y1 z1 
Ae = det 1 y 2 z 2 
1
(3.30)
2
1 y 3 z 3 

41
Each element may now be characterized by the shape functions of Equation

3.28. In vector form this equates to the following row vector of shape functions for a

single element.

N e = [ N1 N 2 N 3 ] (3.31)

The y and z coordinates of the problem space may now be represented in terms of the

shape functions as:

3
 y1 
y = ∑ N i yi = [ N1 N 2 N 3 ]  y 2  (3.32)
i =1
 y3 

3
 z1 
z =∑ N i z i = [ N1 N 2 N 3 ]  z 2  (3.33)
i =1
 z3 

Integration of elements described by area coordinates is conveniently performed

using the following area integral relation [9, 14]:

a ! b ! c!
∫N N 2b N 3c dAe = 2 Ae
a
1 (3.34)
Ae
(a + b + c + 2) !

This equation will be used in Section 3.3 during the finite element formulation of the

problem of torsion and cross section properties for the 3-node triangle.

3.2.2. Six Node Triangle: Isoparametric Coordinates

In addition to the straight-sided 3-node element, the 6-node curved-sided

element is desired. The incorporation of a 6-node element will be especially usefully in

application towards cross sections having curved boundaries. In this case, fewer

42
elements will be needed to accurately capture the curvature of the surface. See Figure

3.2.

Figure 3.2: Representative six node triangle divided with curved sides

The curved-sided 6-node element is developed from Reference [20] and [21].

The shape function vector describing the displacement behavior of all six nodes is given

as:

N e = [ N1 N 2 N 3 N 4 N 5 N 6 ] (3.35)

where

N1 = (1 − (ξ + η )) (1 − 2(ξ + η )) (3.36)

N 2 = ξ (2ξ − 1) (3.37)

N 3 =η (2η − 1) (3.38)

N 4 = 4ξ (1 − (ξ + η )) (3.39)

N 5 = 4ξη (3.40)

N 6 = 4η (1 − (ξ + η )) (3.41)

43
With regard to the actual element, N1, N2, and N3 are the shape functions for the three

corner nodes, moving in counter-clockwise order. N4, N5, and N6 are the shape

functions for the midside nodes such that N4 defines the midside node between N1 and

N2, N5 defines the midside node between N2 and N3, and N6 defines the midside node

between N3 and N1.

The y and z coordinates of the problem space may now be represented in terms

of the shape functions as:

 y1 
y 
 2
6 y 
y = ∑ N i yi = [ N1 N 2 N 3 N 4 N5 N6 ]  3  (3.42)
i =1  y4 
 y5 
 
 y 6 

 z1 
z 
 2
6 z 
z =∑ N i z i = [ N1 N 2 N 3 N 4 N5 N6 ]  3  (3.43)
i =1 z4 
 z5 
 
 z 6 

Integration of the 6-node curved –sided element described by isoparametric

coordinates is performed numerically using Gauss Quadrature. For the integration of a

given function I (ξ ,η ) , the following integration approximation from Reference [20]

and [21] is applied:

1 1−ξ
1 n
∫ ∫ I (ξ ,η ) J e dη dξ = ∑ wi I (ξ i ,η i ) J ei
2 i
(3.44)
0 0

44
where n is the number of Gauss points used, wi, ξ i , and η i are the weighting function

and shape functions at each Gauss point. J e is the determent of the Jacobian matrix,

Je, which is defined in Equation 3.53.

For a 6-node triangle with n equals 6 Gauss points, the following table of values

for weighting functions and shape functions at each point is used in Equation 3.44.

Table 3.1: Weighting function and shape function values for each Gauss point [20, 21]

Equation 3.44 will be used in Section 3.3 during the finite element formulation of the

problem of torsion and cross section properties for the 3-node triangle.

3.3 FE Formulation of Discretized Problem (A u = b)

3.3.1. Left Hand Side: Stiffness Matrix

The Ritz-Galerkin discretization obtained in Equation 3.20 contains several

important contributions to the finite element formulation of the problem of torsion.

Equation 3.22 illustrates one such contribution: the element stiffness matrix.

K e = ∫ ∇N k ⋅ ∇N Tj dA (3.45)
Ae

where N represents the [1 x n] shape function vector defining an element of n-nodes

∂ ^ ∂ ^
and ∇ = j+ k .
∂y ∂z

45
3.3.1.1. Three Node Triangles

In order to develop the stiffness matrix representation for the three node element

defined in Section 3.2.1, it is first necessary to describe the partial derivatives of the

shape function over the element as seen in Equation 3.44. For a three node triangle this

will result in the following two equations.

∂N 1  ∂N1 ∂N 2 ∂N 3  1
= = [b1 b2 b3 ] (3.46)
∂y 2 Ae  ∂y ∂y 
∂y  2 Ae

∂N 1  ∂N1 ∂N 2 ∂N 3  1
= = [c1 c2 c3 ] (3.47)
∂z 2 Ae  ∂z ∂z 
∂z  2 Ae

with Ae , ci, and bi defined by Equations 3.29 and 3.30.

Equation 3.45 may now be rewritten in terms of these derivatives as

 ∂N T ∂N T  ∂N ∂N 
Ke = ∫  j+ k  j+ k  dAe (3.48)
 ∂y
Ae 
∂z  ∂y ∂z 

and following multiplication of like terms becomes [5, 14]

 ∂N T ∂N ∂N T ∂N 
Ke = ∫  +  dAe (3.49)
 ∂y ∂y ∂z ∂z 
Ae  

∂N ∂N ∂N T ∂N T
where and are row vectors (1 x 3) and the transposed and are
∂y ∂z ∂y ∂z

column vectors (3 x 1) resulting in a (3 x 3) stiffness matrix. Substituting Equations

3.46 and 3.47 into Equation 3.49 obtains

  b1   c1  
 1 b [b b 1   
Ke = ∫ 2  2 1 2 b3 ] + c2 [c1 c2 c3 ] dAe (3.50)
Ae  4 Ae 4 Ae2   
 b3  c3  

46
Since everything in this equation is known via Equations 3.29 and 3.30 for the given y

and z coordinates of each node for each three node element, the area integral applies

only to the differential operator, dAe. The result, after some minor simplification, takes

the form:

b12 + c12 b1b2 + c1c 2 b1b3 + c1c3 


1  
Ke = ... b22 + c22 b2 b3 + c2 c3  (3.51)
4 Ae 
 .... ... b32 + c32 

Finally, Equation 3.51 represents the finite element formulation for the element stiffness

matrix of a three node triangle. The result is a symmetric matrix of dimension [3 x 3].

A global stiffness matrix is then assembled using the direct stiffness method [14].

3.3.1.2. Six Node Triangles

In order to develop the stiffness matrix representation for the 6-node curved-

sided element defined in Section 3.2.2, it is first necessary to expand Equation 3.45 as

before:

 ∂N T ∂N ∂N T ∂N 
K e = ∫ ∇N k ⋅ ∇N dA = ∫
T  +  dAe (3.52)
j  ∂y ∂y ∂z ∂z 
Ae Ae  

where N represents the row vector of the shape function given by Equation 3.34.

Rewriting Equation 3.52 by factoring out the non-transposed vectors leads to the

convenient representation of:

 ∂N 
 ∂N T ∂N T   ∂y 
Ke = ∫    dAe (3.53)
 ∂y
Ae 
∂z   ∂N 
 ∂z 

47
Reference [5] defines the follow with respect to the above equation:

Ae = ∫
Ae
dAe = ∫ J e dη dξ (3.54)

 ∂y ∂z 
 ∂ξ ∂ξ 
with J e =   (3.55)
 ∂y ∂z 
 ∂η ∂η 

 ∂N   ∂N ∂η ∂N ∂ξ   ∂η ∂ξ   ∂N   ∂N 
   ∂η +
∂y ∂ξ ∂y   ∂y ∂y   ∂η  = J −1  ∂η  = B
 
 ∂y = =
∂η ∂N ∂ξ   ∂η 
and (3.56)
 ∂N   ∂N ∂ξ   ∂N  e
 ∂N 
+
 ∂z   ∂η ∂z ∂ξ ∂z   ∂z ∂z   ∂ξ   ∂ξ 

where the differential area dAe is defined in terms of isoparametric operators, Je is a

matrix of derivatives of y and z with respect to the shape function variables, ξ i and η i

called the Jacobian matrix, and a new variable is introduced; the B matrix. The B

matrix, as seen above, is a (2x6) matrix composed of the product of the shape function

derivatives and the inverse of the Jacobian matrix.

Likewise Reference [5] defines:

 ∂N T ∂N T 
=B
T
 (3.57)
 ∂y ∂z 

which is a (6x2) matrix of shape function derivatives.

These relations provide a new form for element stiffness matrix of Equation 3.53:

1 1−ξ

Ke = ∫ ∫ Be Be J e dη dξ
T
(3.58)
0 0

Integration of Equation 3.58 is performed through application of the Gauss

Quadrature methods outlined in Subsection 3.2.2 by Equation 3.44 and Table 3.1. If
48
I (ξ ,η ) = Be Be
T
(3.59)

then the element stiffness matrix of Equation 3.56 is approximated by:

1 1−ξ
1 n
ke = ∫ ∫ Be Be J e dη dξ = ∑
T T
wi Bei Bei J ei (3.60)
0 0
2 i

where n is the number of Gauss points used, wi, ξ i , and η i are the weighting function

and shape functions at each Gauss point given by Table 3.1. The element stiffness

matrix has dimensions of (6x6). A global stiffness matrix is then assembled using the

direct stiffness method [14].

3.3.2. Right Hand Side: Volume Force and Boundary Conditions

Now that the left hand side of the Ritz-Galerkin discretization of Equation 3.17

has been addressed, attention may be turned to the right hand side which contains three

important contributions to the finite element formulation for the problem of torsion.

n
b= ∫ f N T dAe + ∫ g N T dΓN − ∑U k ∫ ∇N T ∇N dAe (3.61)
Ae ΓN k =1 Ae

The first integral of Equation 3.61 relates to the internal volume force for a

given cross section. The second integral accounts for the necessary Neumann boundary

conditions placed on the cross section, while the third integral accounts for the Dirichlet

boundary conditions which are required to “pin down” the stiffness matrix to obtain a

non-singular system [11, 16].

3.3.2.1. Three Node Triangles

The first integral of Equation 3.61, as discussed at the beginning of Chapter 3,

vanishes in the warping function formulation of the torsion problem because the

49
function f is taken to be zero. This can be seen by examining Equations 2.40 and 3.1.

However, in the case of the stress function formulation, f is equal to 2; see Equation

2.60 and 3.1.

For the case of the stress function formulation, as developed in Chapter 2, the

first integral of Equation 3.61, shown below,

b= ∫fN dAe + ....


T
(3.62)
Ae

with N T = [N1 N 3 ] is solved using the following area integral relation [9, 14]:
T
N2

a ! b ! c!
∫N N 2b N 3c dAe = 2 Ae
a
1 (3.63)
Ae
(a + b + c + 2) !

Applying this to Equation 3.62 provides a numerical solution for the integral of the

volume force over a triangular, 3-node element [11, 14].

 N1  1
 N  dA = f Ae 1
∫fN dAe = f ∫
T
 2 e (3.64)
3 
Ae Ae
 N 3  1

where f depends on the problem; warping function, f=0, or stress function, f=2.

The second integral of Equation 3.61 defines how the Neumann boundary

condition is applied to the problem of torsion.

b = .... + ∫ g N T d ΓN +.... (3.65)


ΓN

50
∂ω
Here the function g represents the warping normal derivative, , found in the warping
∂n

formulation of the torsion problem and N T = [N1 N 3 ] is the column vector of


T
N2

element shape functions.

For the line integral of Equation 3.65, it is useful to employ the following line

integral relation from reference [14]:

∫N ( s) N ( s) ds = ∫ l1a l 2b Le dl2
a b
1 2 (3.66)
L 0

where N1 = l1 , N 2 = l 2 , s = Le l 2 , and ds = Le dl2 are obtained from the natural

coordinate relations for the one-dimensional case of a line element. Through

application of this integration relation, Equation 3.65 becomes [11, 14]:

1! L
∫gN d s = g Le = eg
T
(3.67)
L
(1 + 1) ! 2

where Le is the length of the line defined by the node pair of an element that is touching

a boundary and g is the warping normal derivative given by Equation 2.45.

∂ω  dz   − dy 
= zn y − ynz = z   − y  (3.68)
∂n  ds   ds 

where z and y correspond to the y- and z-coordinate location of the boundary condition

 dz   − dy 
and   and   represent the boundary normal vector components ny and nz.
 ds   ds 

The third and final portion of Equation 3.59 defines how the Dirichlet boundary

conditions are applied to the problem of torsion for both the warping and stress function

formulations and is shown below in Equation 3.69.

51
n
b = ... − ∑ U k ∫ ∇N T ∇N dAe (3.69)
k =1 Ae

There are two distinct pieces that can be seen in the above equation. The first portion is

a column vector containing all of the nodal values of U that are known explicitly

through Dirichlet boundary conditions for the given cross section. These values will be

called UD. The second portion of Equation 3.69 is exactly the same as the stiffness

matrix relation discussed in the previous section, Equation 3.51.

The effect of Equation 3.69 can best be demonstrated through a formal

explanation of Dirichlet boundary condition on the linear set of equations (A x = b) as

presented in reference [9].

 K 11 K 12   xu  bD 
= (3.70)
K
 21 K 22   x D   bu 

The subscript D refers to known values and the subscript u refers to unknown values.

The above example represents two linear equations, the first of which, after

multiplication of the x-vector into the stiffness matrix, K, yields:

K 11 ⋅ xu + K 12 ⋅ x D = bD (3.71)

Ignoring the second equation resulting from Equation 3.48 and subtracting the known

values of Equation 3.71 over to the other side of the equals sign gives:

K 11 ⋅ xu = bD − K 12 ⋅ x D (3.72)

which translates to

A ⋅ x (unknown ) = b − A ⋅ x ( known ) (3.73)

52
such that known values of the displacement vector, x, are subtracted from the volume

force.

3.3.2.2. Six Node Triangles

It is necessary only to address the first integral of Equation 3.61 representing the

element volume force when applied to the 6-node triangle. The second and third

integrals relating the Neumann and Dirichlet boundary conditions back to the volume

force remain the same for the 6-node triangle as they were for the 3-node triangle in

Equations 3.67 and 3.72.

The element volume force is again represented by the following equation:

b= ∫fN
T
dAe (3.75)
Ae

This integral must now be integrated using the Gauss Quadrature integration as

used for the element stiffness matrix. By applying Equation 3.75 to the Gauss

Quadrature integration equation, the following approximate relation is obtained for the

element volume force, be:

1 1−ξ n
1
be = f ∫ ∫ N T (ξ ,η ) J e dη dξ =
2
f ∑w N
i
i
T
(ξ i ,η i ) J ei (3.76)
0 0

where n is the number of Gauss points used, wi, ξ i , and η i are the weighting function

and shape functions at each Gauss point given by Table 3.1. Here again, f is a function

of the problem being solved; either the stress function or warping function.

53
3.4 Beam Cross Section Properties

This section presents the finite element formulations for the cross section

property equations of outlined in Chapter 2 for warping independent and warping

dependent cross section properties.

3.4.1. Warping Independent Cross Section Properties

The warping independent cross section properties provided in Section 2.1 are

rederived for finite element application in the following subsections for both three node

triangles and six node triangles.

3.4.1.1. Three Node Triangles

All of the following warping independent cross section properties that require

integration, with the exception of area, were integrated using Equation 3.34 and is

repeated here in Equation 3.77.

a ! b ! c!
∫N N 2b N 3c dAe = 2 Ae
a
1 (3.77)
Ae
(a + b + c + 2) !

Total area for a given cross section is obtained through the summation of the

following elemental area equation that was previously developed in the three node

triangle section of Section 3.2.1.

1 y1 z2 
1 
Ae = det 1 y 2 z 2  (3.78)
2
1 y3 z3 

The first moments of the area, Qy and Qz, for a given cross section are defined

by the following area integrals as seen in section 2.1.

54
Q y = ∫ z dA (3.79)
A

Q z = ∫ y dA (3.80)
A

where y and z may be replaced with the following nodal representations for an element:

 y1   z1 
 
y = [ N ][ y ] = [ N1 N 2 N 3 ]  y 2  , z = [ N ][ z ] = [ N1 N 2 N 3 ]  z 2  (3.81)
 y3   z3 

Insertion of these relations into Equations 3.79 and 3.80 along with the application of

the integral relation for area coordinates given in Equation 3.77 supplies the following:

Ae
Qy e = [z1 + z 2 + z 3 ] (3.82)
3

Ae
Qz e = [ y1 + y2 + y3 ] (3.83)
3

where total values may be obtained for the entire cross section by summing these

equations for each element on the cross section.

The location of the centroid remains the same with:

Qz
yc = (3.84)
A

Qy
zc = (3.85)
A

where Qy and Qz represent the total first moments of the area.

The finite element formulation of the moments of inertia and product of inertia

become:

55
I ye =
Ae 2
6
[
z1 + z 22 + z 32 + z1 z 2 + z 2 z 3 + z1 z 3 ] (3.86)

I ze =
Ae 2
6
[
y1 + y 22 + y32 + y1 y 2 + y 2 y3 + y1 y3 ] (3.87)

Ae
I y ze = [ y1 z1 + y 2 z 2 + y3 z 3 ] + Ae [ y1 z 2 + y2 z1 + y1 z 3 + y3 z1 + y 2 z3 + y3 z 2 ] (3.88)
6 12

with total values for these three properties being, again, the sum of each over every

element on the domain.

3.4.1.2. Six Node Triangles

All of the following warping independent cross section properties that require

integration were integrated using Gauss Quadrature numerical integration and Table

3.1. For the Gauss Quadrature integration of a given function I (ξ ,η ) , the following

integration approximation from Reference [20] and [21] is applied:

1 1−ξ
1 n
∫ ∫ I (ξ ,η ) J e dη dξ = ∑ wi I (ξ i ,η i ) J ei
2 i
(3.89)
0 0

where n is the number of Gauss points used for the element, wi, ξ i , and η i are the

weighting function and shape functions at each Gauss point. J e is the determent of the

Jacobian matrix, Je. The Jacobian matrix as given in Equation 3.55, repeated below:

 ∂y ∂z 
 ∂ξ ∂ξ 
Je =   (3.90)
 ∂y ∂z 
 ∂η ∂η 

Element area for a six node triangle is obtained using Equation 3.73 [5].

56
1 1−ξ

Ae = ∫
Ae
dAe = ∫
0

0
J e dη dξ (3.91)

Applying Gauss Quadrature integration yields:

1 n
Ae = ∑ wi J ei
2 i
(3.92)

When summed over all six Gauss points, the element solution is reached. Through

summation of Equation 3.92, the total area for a cross section may be obtained.

The first moments of the area, Qy and Qz, for a given cross section are defined

by the following area integrals as seen in section 2.1.

Q y = ∫ z dA (3.93)
A

Q z = ∫ y dA (3.94)
A

where y and z may be replaced with the following nodal representations for an element

with its shape functions:

6
ye = ∑ N i yi = [N ][ y ] (3.95)
i =1

6
z e = ∑ N i z i = [N ][z ] (3.96)
i =1

[N] is the (1x6) row vector of shape functions and [y], [z] are the (6x1) column vectors

of node y- and z-coordinates for each element.

Inserting Equations 3.95 and 3.96 into Equations 3.93 and 3.94 respectively, and

applying the Gauss Quadrature integration method yields:

57
1 1−ξ
1 n
Q ye = ∫
0
∫ [N ][z ] J e
0
dη dξ = ∑ wi [N i ][zi ] J ei
2 i
(3.97)

1 1−ξ n
1
Q ze = ∫
0
∫ [N ][y ] J e dη dξ = 2 ∑i wi [N i ][ yi ] J ei
0
(3.98)

The location of the centroid remains the same with:

Qz
yc = (3.99)
A

Qy
zc = (3.100)
A

where Qy and Qz represent the total first moments of the area.

The finite element formulation of the moments of inertia and product of inertia

become:

1 1−ξ
1 n
I ye = ∫ ∫ ( [N ][z ] ) J e dη dξ = ∑ wi ( [N i ][z i ] ) J ei
2 2
(3.101)
0 0
2 i

1 1−ξ
1 n
I ze = ∫ ∫ ( [N ][ y] ) J e dη dξ = ∑ wi ( [N i ][ y i ] ) J ei
2 2
(3.102)
0 0
2 i

1 1−ξ
1 n
I y ze = ∫
0
∫ ( [N ][y] )( [N ][z ] ) J e
0
dη dξ = ∑ wi ( [N i ][yi ] )( [N i ][z i ] ) J ei (3.103)
2 i

3.4.2. Warping Dependent Cross Section Properties

The warping dependent cross section properties provided in Section 2.2 are

rederived toward a finite element application in the following subsections for both three

node triangles and six node triangles.

58
3.4.2.1. Three Node Triangles

All of the following warping dependent cross section properties were integrated

using the same integral equation that was used for the warping independent cross

section properties for three node triangles and is repeated here in Equation 3.104 [14].

a ! b ! c!
∫N N 2b N 3c dAe = 2 Ae
a
1 (3.104)
Ae
( a + b + c + 2) !

The first moment of warping, as given in Section 2.2.2, is specified as the

following area integral:

Qω e = ∫ ω dA (3.105)
Ae

where the warping function, ω, may be represented as below for each element.

3
ω1 
ω = ∑ N i ω i = [ N 1 N 2 N 3 ] ω 2  (3.106)
i =1
ω 3 

Applying Equation 3.104 as previously, following the insertion of Equation 3.73, the

first moment of warping for each element becomes

Ae
Qω e = [ω1 + ω2 + ω3 ] (3.107)
3

The Sectorial products of inertia are likewise integrated in such a manner,

producing Equations 3.75 thru 3.77 below.

Iω e =
Ae 2
6
[
ω1 + ω 22 + ω32 + ω1ω 2 + ω 2ω3 + ω1ω3 ] (3.108)

59
Ae
I yω e = [ y1ω1 + y 2ω2 + y3ω3 ] + Ae [ y1ω 2 + y 2ω1 + y1ω3 + y3ω1 + y 2ω3 + y3ω 2 ] (3.109)
6 12

Ae
I zω e = [z1ω1 + z 2ω2 + z3ω3 ] + Ae [z1ω2 + z 2ω1 + z1ω3 + z3ω1 + z 2ω3 + z3ω2 ] (3.110)
6 12

Equations 3.108 – 3.110 may be summed across every element on the cross

section to obtain the total cross section value for each.

The finite element formulation of the torsional constant, J, for the warping

function approach is given by the following equation from reference [5].

J = I y + I z − bT ω (3.111)

Where bT is the row vector, (1 x n), representing the global volume force vector and ω is

the column vector, (n x 1), representing the global warping function vector

The finite element formulation of the torsional constant, J, for the stress

function approach is given by the following equation from reference [5].

J e = 2 ∫ψ dA (3.112)
Ae

Following area coordinate integrations:

3
2 Ae
Je =
3
∑ [ψ
1
1 +ψ 2 + ψ 3 ] (3.113)

The warping constant, Γ , remains the same as in Section 2.2.2 with:

2
Q
Γ = I ω − ω − y s2 I y + 2 y s z s I yz − z s2 I z (3.114)
A

where ys and zs are the coordinates for the shear center whose equations are found in

Section 2.2.2.

60
3.4.2.2. Six Node Triangles

The warping dependent cross section properties for the 6-node triangle after

applying the Gauss Quadrature take the following form:

The first moment of warping is given as Equation 3.115 below.

Qω e = ∫ ω dA (3.115)
Ae

where the warping function, ω, may be represented as below for each element.

6
ω = ∑ N i ω i = [ N ][ω ] (3.116)
i =1

[N] is the (1x6) row vector of shape functions and [�] is the (6x1) column vectors of

nodal warping function values for each element.

Equation 3.82 becomes:

1 1−ξ n
1
Qωe = ∫
0
∫ [N ][ω ] J e dη dξ = 2 ∑i wi [N i ][ωi ] J ei
0
(3.117)

The Sectorial products of inertia are likewise integrated in such a manner,

producing Equations 3.118 – 3.120 below.

1 1−ξ
1 n
I ωe = ∫ ∫ ( [N ][ω ] ) J e dη dξ = ∑ wi ( [N i ][ω i ] ) J ei
2 2
(3.118)
0 0
2 i

1 1−ξ
1 n
I yωe = ∫
0
∫ ( [N ][y] )( [N ][ω ] ) J e
0
dη dξ = ∑ wi ( [N i ][yi ] )( [N i ][ωi ] ) J ei (3.119)
2 i

1 1−ξ
1 n
I zωe = ∫
0
∫ ( [N ][z ] )( [N ][ω ] ) J e
0
dη dξ = ∑ wi ( [N i ][zi ] )( [N i ][ωi ] ) J ei
2 i
(3.120)

61
Equations 3.118 – 3.120 may be summed across every element on the cross

section to obtain the total cross section value for each.

The finite element formulation of the torsional constant, J, for the warping

function approach stays the same that used for the 3-node triangles [5].

J = I y + I z − bT ω (3.121)

The finite element formulation of the torsional constant, J, for the stress

function approach however must be integrated over the 6-node element.

Applying Gauss Quadrature integration methods provide:

1 1−ξ
1 n
Je = ∫
0
∫ [N ][ψ ] J e
0
dη dξ = ∑ wi [N i ][ψ i ] J ei
2 i
(3.122)

The warping constant, Γ , remains the same as in Section 2.2.2 with:

2

Γ = Iω − − y s2 I y + 2 y s z s I yz − z s2 I z (3.123)
A

where ys and zs are the coordinates for the shear center whose equations are found in

Section 2.2.2.

62
CHAPTER 4

RESULTING CROSS SECTION PROPERTY FEM TOOL

Having now established the finite element formulation for the problem of

torsion and all of the desired cross section properties, it is necessary to translate these

aspects into an appropriate programming environment. For this project, the MATLAB

programming environment has been chosen. MATLAB, which stands for matrix

laboratory, is an environment that utilizes arrays as its basic element. No dimensioning

is necessary with the array element in this system, which makes it very usefully for

solving systems of matrices and vectors [17, 18]. Other programming languages such

as FORTRAN, while incredibly powerful, only use scalar elements requiring more

complicated programming techniques to represent problems that rely heavily on

matrices [17].

As outlined in Chapter 1, the purpose of this project is to create a finite element

tool for the determination of beam cross section properties in the case of pure torsion

using the MATLAB programming environment. The objective of this FEM tool is to

provide a capability for the calculation of these properties for cross sections of arbitrary

shape, with singly or multiply connected domains, of either open or closed

configuration. To accomplish these objectives, the finite element formulations of

Chapter 3 have been implemented in the MATLAB programming environment with the

end result being a user friendly, intuitive tool for calculating section properties.

63
4.1 Overview of FEM Code- Basic Structure

The finite element tool for calculating cross section properties presented here is

based in part on the finite element code, FEM_50, which provides a generic framework

for the solution of the Laplace and Poisson equations using 3-node triangles and 4-node

quadrilateral elements [11]. The FEM_50 code has been modified and expanded in

order to solve the problem of torsion for the warping function and each of the desired

cross section properties using both 3-node, straight edged triangles and 6-Node, curved-

sided triangles.

The main body of the code may be divided generically into three sections: input,

analysis, and output; each of which is discussed in detail in the remaining sub-sections

of this chapter. A brief overview of the tool is presented here. See Figure 4.1.

Before analysis begins, it is first necessary to construct proper input files that

define the geometry of the cross section. This is accomplished by first meshing the area

of interest to produce a collection of nodes and elements over given surface and then

linking this data to the analysis tool in the proper format. There are three types of input

files required for this FEM tool, and all may be constructed using the meshing software,

EasyMesh [19], or in the case of the 6-Node triangles, by hand. One input file contains

nodal y- and z-coordinates, one contains element-node data, and the input file type

identifies node pairs corresponding to Neumann and/or Dirichlet type boundary

conditions (coordinates.dat, elements3.dat, elements6.dat, neumann.dat, and

dirichlet.dat).

64
Figure 4.1: Basic logic flow for cross section FEM Tool

The analysis portion of the tool consists of one main driver program (BCSP.m)

and five external functions (f.m, u_d.m, stima3.m, stima6.m, and volume6.m).

Depending on the type of problem being solved, warping function or stress function,

each of the external functions listed may or may not be used. The input data is read into

the main program and used to solve for either the warping function or the stress

function. If the warping function approach is chosen then both warping independent

and warping dependent cross section properties are calculated (Sections 2.1 and 2.2). If

the stress function is used, then only the warping independent cross section properties

and the torsional constant, J, are calculated. Figure 4.1 illustrates the project flow as

described above.

65
4.2 Required Inputs

The beam cross section property finite element tool, or BCSP, requires data

given by three types of input files in order to operate. As mention in Section 4.1 these

input files contain:

1) Nodal Coordinates (coordinates.dat)

2) Element Data (elements3.dat, elemtents6.dat)

3) Boundary Condition Data (neumann.dat, dirichlet.dat)

To obtain the data above, the cross section of interest is discretized, or meshed,

into many small triangular elements composed of three or six nodes, depending on the

user’s preference. Meshing a cross section with 3-Node triangles is performed using

EasyMesh, a GNU meshing tool developed by Reference 19. Meshing a cross section

with 6-Node curved-sided triangles, however, has been performed by hand since access

to a meshing tool for this type of element is not readily available for this current project.

4.2.1. EasyMesh Output Files

EasyMesh version 1.4 has been used to mesh 3-Node straight edge triangles for

this project. The meshing tool, when given an input file that defines the boundaries of a

certain cross section, provides three output files containing nodal coordinates, element

node numbers, and side node numbers. Instructions and examples for input file creation

and operation of EasyMesh may be found on the tool’s website [19].

After running the mesher for a specific cross section, EasyMesh provides the

three outputs listed above. The output containing nodal coordinate data has the format

shown below [19].

66
• First line: (number of nodes)

• Following lines: (node #) (y-coordinate) (z-coordinate) (marker)

The marker entry in the output above is used to identify nodes lying in regions or on

boundaries of interest that are set by the user in the EasyMesh input file. The nodal

output file is saved as filename.N when meshing is complete.

The second output from EasyMesh contains the element node data and has the

following format [19]:

• First line: (number of elements)

• Following lines:

(element #) (i) (j) (k) (ei) (ej) (ek) (si) (sj) (sk) (xv) (yv) (marker)

Here i, j, and k represent the nodes that belong to each element and are listed in counter-

clockwise order. For the element output file, (marker) is used to denote locations

composed of different materials. The rest of the data listed in this output file provides

detailed information about surrounding elements, adjoining sides, and element

circumcenter coordinates. The element output file is saved as filename.E when meshing

is complete.

The third output from EasyMesh contains the element side node data and has the

following format [19]:

• First line: (number of sides)

• Following lines: (side #) (c) (d) (ea) (eb) (marker)

where (c) and (d) indicate the start and end points of the side, (ea) and (eb) represent the

element to right or left of the side, and (marker) can be used to identify sides lying in

67
regions or on boundaries of interest that are set by the user in the EasyMesh input file.

The side output file is saved as filename.S once meshing is complete.

When each of the three EasyMesh output files have been obtained for a specific

cross section it is next important to extract only the necessary data required to run the

finite element tool, BCSP. Since a given problem has the potential to have a large

number of nodes, elements, and sides, it is convenient to automate the sorting process in

order to gather only the needed data, in the proper format, in a short amount of time.

This has been done through the creation of a FORTRAN code and is discussed below in

Section 4.2.1.

4.2.2. BCSP Finite Element Tool Input Files

As mentioned in the introduction to Section 4.3 above, the BCSP code requires

three types of input files to provide the needed data defining a cross section. The first

input file contains only the nodal y- and z-coordinates for each node. The input file

format for each row of data is simply:

• (node #) (y-coordinate) (z-coordinate)

The node coordinate input file must be a data file (.dat) to be accepted by MATLAB

and must be named coordinates.dat to be accepted by the BCSP code.

The second input file type contains the element node data for each element on

the cross section. The element input file format for each row of data is as follows for

the 3-Node triangle and then the 6-Node triangle:

• (element # ) (node 1) (node 2) (node 3)

• (element #) (node 1) (node 2) (node 3) (node 4) (node 5) (node 6)

68
The 3-Node triangle nodes must be listed in a counter-clockwise fashion to insure

proper results as dictated by the area coordinate development of Chapter 3. The 6-Node

triangle format requires that node 1, node 2, and node 3 be comprised of the nodes

occupying the vertices of the triangle and are listed in a counter-clockwise direction.

The midside nodes must then be defined as node 4, node 5, and node 6 listed in a

counter-clockwise direction as well, with node 4 being the midside node between nodes

1 and 2, node 5 being the midside node between nodes 2 and 3, and finally node 6 being

the midside node between nodes 3 and 1. The 3-node and 6-node element input files

must be data files (.dat) to be accepted by MATLAB and must be named elements3.dat

and elements6.dat to be accepted by the BCSP code.

The third input file type contains the data required for application of boundary

conditions. The two possible boundary conditions for this project are of the Neumann

and Dirichlet type. In each case, the input file should contain the node pairs that define

every location where boundary conditions will be applied. The boundary condition

input file format for each row of data is as follows:

• (side #) (node 1) (node 2)

The input files containing data corresponding to Neumann boundary conditions should

be named neumann.dat and input files containing data corresponding to Dirichlet

boundary conditions should be named dirichlet.dat.

The number and type of the above inputs depends on the problem being solved

in the finite element tool. If the user chooses the stress function approach, only

Dirichlet boundary conditions will be used. If the warping function approach is used,

69
then Neumann boundary conditions must also be applied. The type of element used can

be either 3-node or 6-node depending on the user’s preference. In all cases, it is

imperative that every input files be present in the folder with the main driver

program BCSP, whether they are used or not, to ensure functionality (i.e.

coordinates.dat, elements3.dat, elemtents6.dat, neumann.dat, and dirichlet.dat).

4.2.3. BCSP Input File Creator

There is an obvious difference between the format of the output files obtained

from EasyMesh and the required input format for the BCSP code. To remedy this a

FORTRAN code was produced that can quickly collect only the pertinent data from the

EasyMesh output files for the case of the 3-node triangle, returning four files of the

proper format, file name, and file type required by MATLAB and the BCSP code. A

copy of this FORTRAN code may be found in Appendix B.

Given the format of the EasyMesh output files filename.N, filename.E, and

filename.S which represent the nodal, elemental, and side data of the meshed area, the

FORTRAN input file creator must read only the data corresponding to the BCSP input

file format.

For the EasyMesh nodal output file:

• First line: (number of nodes)

• Following lines: (node #) (y-coordinate) (z-coordinate) (marker)

only the y- and z-coordinates on each row are selected by the FORTRAN code and

written into a data file (.dat) named coordinates.dat.

For the EasyMesh element output file:

70
• First line: (number of elements)

• Following lines:

(element #) (i) (j) (k) (ei) (ej) (ek) (si) (sj) (sk) (xv) (yv) (marker)

only the i, j, and k node numbers on each row are selected by the FORTRAN code and

written into a data file (.dat) named elements3.dat.

For the EasyMesh side output file:

• First line: (number of sides)

• Following lines: (side #) (c) (d) (ea) (eb) (marker)

the FORTRAN input file creator must identify and select only the sides (node pairs) that

the user has marked via the marker option, in EasyMesh, as lying on boundaries where

B.C.’s are to be applied. For the sides located along the “marked” boundaries, only the

c and d node numbers representing the side (node pair) on each row are selected by the

FORTRAN code and written into a data file (.dat) named neumann.dat and

dirichlet.dat. If a specific node pair lies on a Dirichlet boundary it should be designated

by a marker value of 1 in EasyMesh. If the node pair lies on a Neumann boundary it

should be designated by a marker value of 3 in EasyMesh.

4.3 Analysis and Output

In general, the finite element tool, BCSP.m, presented here solves firstly the set

of linear equations derived in Chapter 3 as

Au=b (4.1)

for a given cross section where the left side of the equation contains A, the stiffness

matrix, and u, the stress or warping function. The right side of the equation contains b,

71
the volume force and the related Neumann and Dirichlet boundary condition relations

that apply. When the stress function or warping function has been found, BCSP.m then

calculates the warping independent and warping dependent cross section properties for

the cross section. Throughout this process, as illustrated in Figure 4.2, six external

functions are called by the driver program to aid in calculations for the stiffness matrix,

A and the volume force, b. These functions are detailed in Subsection 4.3.1 during a

discussion of BCSP.m. Subsection 4.3.2 presents a brief description on how the code

may be conditioned to solve for either the stress function or the warping function.

Figure 4.2: Main driver program BCSP.m and external functions

4.3.1. BCSP Finite Element Tool

A walkthrough of the driver program, BCSP.m, and the six external functions

seen in Figure 4.2 above, is presented here. The BCSP code is divided among the

following five subsections, 4.3.1.1 - 4.3.1.5.

4.3.1.1. Read Inputs and Symmetry Condition

The first action performed by the driver program is to read each of the input

files necessary for analysis using the load command. This action is performed on all
72
five input files (coordinates.dat, elements3.dat, elemtents6.dat, neumann.dat, and

dirichlet.dat) even when empty.

Figure 4.3: Input file read-in

Next, in Figure 4.4, a symmetry condition is introduced allowing the user to take

advantage of symmetry, modeling only that portion of the geometry. When BCSP is

executed, the user is prompted to input a value for the symmetry condition being used.

If full geometry of the cross section is defined by the input files, zero is selected

indicating zero symmetry. If ½ symmetry exists about the y-axis for the cross section, 1

is selected. Similarly, if ½ symmetry exists about the z-axis for the cross section, 2 is

selected. Finally, if the geometry is symmetric about the y- and z-axes, the user may

model 1/4th geometry and select a symmetry value of 4.

73
Figure 4.4: Cross section symmetry factor

4.3.1.2. Stiffness Matrix Development

Having loaded all node, element, and boundary condition data from each

input file, the BCSP code’s next objective is to develop the element stiffness and then

global stiffness matrices for the problem of torsion. Figure 4.5 is the main portion of

the external function stima3.m whose purpose is to calculate the element stiffness

matrix for every 3-node element given in the elements3.dat input file. The logic for the

stima3.m function was developed by Reference [11] and is equivalent to the 3-node

stiffness matrix formulation shown in Equation 3.51 of Chapter 3.

Figure 4.5: 3-Node triangle element stiffness matrix (stima3.m) [11]

Figure 4.6 and 4.7 show the external function stima6.m, whose purpose is to

calculate the element stiffness matrix for every 6-node element given in the

elements6.dat input file. As discussed in Subsection 3.3.1.2, the integral required for

the determination of the element stiffness matrix is performed using Gauss Quadrature.

Figure 4.6 contains the portion of the function that defines the y- and z- coordinates for

74
all 6-nodes of the element as received from the driver program. The necessary Gauss

points for a 6-node element are also specified here as found in Table 3.1.

Figure 4.6: 6-Node element stiffness matrix (stima6.m)

In Figure 4.7, the Jacobian matrix, J, and the shape functions are used to obtain

the B matrix and the product BTB. This data is obtained for each Gauss point for the

element coordinates given. As derived in Chapter 3 in Equation 3.58, the element

stiffness matrix is given by the summation of the product of BTB and the determinant of

the Jacobian matrix, Je over each Gauss point.

1 n
ke = ∑
2 i
T
wi Bei Bei J ei (4.2)

This result is sent to the main driver BCSP.m after all six Gauss points have been

evaluated.

75
Figure 4.7: 6-Node element stiffness matrix (stima6.m) (continued)

Figure 4.8 depicts the global stiffness matrix development for the 3-Node

triangle and then the 6-Node triangle. Initially a sparse matrix A is created whose size

is (# of nodes by # of nodes). BCSP sends element node data to the stima3.m and/or the

stima6.m functions. Upon return from the external functions, the sparse, empty, matrix

A is filled with element values for the stiffness matrix. The global matrix is filled by

way of the direct stiffness method [14] and is illustrated in Figure 4.8. The for loop

sends every element in turn to the one of the stima functions and upon return, adds the

individual coefficients of the element stiffness matrices accordingly.

76
Figure 4.8: Global stiffness matrix development for 3-Node and 6-Node triangles (BCSP.m)

4.3.1.3. Volume Force Development and Boundary Condition Application

Following the development of the element stiffness matrix is the development

of the element volume force. BCSP first creates a sparse, empty, vector, b whose

dimension is (# of nodes by 1). Next, in the case of volume force for 3-node triangles

(line 70-76, Figure 4.9), element data is used to calculate the volume force of each

element according to Equation 3.62 of Chapter 3 and stored in b according to the node

numbers to obtain the global volume force vector for 3-node triangles adding nodes

together when appropriate. The logic for the element volume force for the 3-node

triangle was developed by Reference [11] according to Equation 3.62 of Chapter 3 and

is conditioned by the external function f.m whose value is zero for the warping function

approach and 2 for the stress function approach.

In the case of 6-node triangles (lines 70-83, Figure 4.9), the main driver must

call the volume6.m function to perform a Gauss Quadrature integral for each element,

77
which is fed back into the main program for addition into the global b vector according

to node number.

Figure 4.9: Global volume force assembly

Figure 4.10 illustrates the application of Neumann boundary conditions to the

problem of torsion. Specifically, this boundary condition applies to the previously

formed global volume force vector, b according to Equations 3.59 and 3.65 of Chapter

3. Line 90 of BCSP checks to see if the neumann.dat boundary condition input file is

empty. If so, the global volume force is left unchanged. When the neumann.dat input

has data within, the global volume force will be augmented by subtracting the warping

normal derivative of Equation 2.45, for each neumann.dat node pair defining the

boundary. Lines 98-100 of Figure 4.10 serve to access every volume force entry that

has a node in common with any neumann.dat nodes and subtract the value of the

warping normal derivative boundary constraint at that node via the external function

warp.m seen in Figure 4.11.

78
Figure 4.10: Neumann boundary condition application

The external function warp.m provides the value of the warping normal

derivative for each pair of nodes sent to it from BCSP. The warping normal derivative

is given in Equation 2.45 of Chapter 2 and is listed here:

∂ω  dz   − dy 
= zn y − ynz = z Ave   − y Ave   (4.3)
∂n  ds   ds 

where zAve and yAve correspond to the midpoint between each node pair given in

 dz   − dy 
neumann.dat and   and   represent the normal vectors ny and nz.
 ds   ds 

Figure 4.11: Warping Normal Derivative (warp.m)

This boundary condition is required for all internal and external boundaries when

solving for the warping function, and is zero when dealing with the stress function.
79
The final type of boundary condition is the Dirichlet boundary condition. It is

applied in BCSP to augment the global volume force and to identify the nodes in the

stiffness matrix that can be disregarded in the solution process of the next section,

Subsection 4.3.1.4. After applying the Dirichlet boundary condition, the global stiffness

matrix will no longer be singular. The theory was presented in Equations 3.68-3.70 of

Chapter 3. Here a sparse, empty vector u is created and filled only the nodes

corresponding to the Dirichlet boundary node pairs given in dirichlet.dat and are each

given a value of zero by the external function u_d.m. The values of u given here

represent either known values of the warping function, ω, or known values of the stress

function, ψ, depending on the problem. The u vector is then multiplied by the cells of

the global stiffness matrix, A, that correspond to the node number and subtracted from

the global volume force vector, b, based on Equation 3.70, as seen in Figure 4.12, line

115 [11].

Figure 4.12: Dirichlet boundary conditions

4.3.1.4. Solution of Linear System: Stress Function/Warping Function

To this point the main driver program, BCSP.m, and its external functions have

developed a global stiffness matrix and a global volume force vector. Boundary on the

problem of torsion have been applied also been applied, resulting in an augmented

80
global volume force and a non-singular global stiffness matrix. Now, the linear system

of equation can be solved for the unknown values of the vector u for either the stress

function or the warping function depending on the problem. Figure 4.13 displays the

manner in which this is achieved in the finite element tool. In the Figure on line 121, a

vector containing all of the free nodes is identified. This vector consists of every node

that is not constrained by a Dirichlet boundary condition, or in other words, all

unknown nodes. Next, a solution for the u vector is obtained by calling only the

coefficients of the stiffness matrix and volume force vector that are attributed to the free

nodes and performing the left inverse of A and b as seen in line 124 of Figure 4.13 [11].

Figure 4.13: Solving the linear system of equations

4.3.1.5. Beam Cross Section Properties

Having solved for the u vector for the stress function or the warping function,

the beam cross section properties may be now calculated. As presented in Chapter 3,

these properties are identified as warping independent and warping dependent. The

BCSP code calculates warping independent and dependent cross section properties first

for 3-node triangles and then for 6-node triangles. Figure 14.14 presents the element

level calculations for the 3-node triangle element. Each cross section property is

implemented here using the area coordinate integral results of Subsection 3.4.1.1.

81
When a value is obtained for each element, it is then summed with the previous

elemental result, producing an answer for the total cross section for each property.

Figure 4.14: 3-Node triangle warping independent and dependent cross section properties

Figure 4.15 illustrate how the y- and z-coordinate of the centroid is found. In

addition to this the parallel axis theorem is applied to the moment of inertia calculations

and the product of inertia to translate them from being defined about the origin to the

centroid location.

Figure 4.15: 3-Node Centroidal calculations and Parallel axis theorem

Figures 4.16 - 4.19 below, demonstrate the calculations performed by BCSP to

calculate the warping independent and dependent beam cross section properties for 6-

Node triangles. It is necessary once again to utilize Gauss Quadrature to calculate the
82
integrals governing the cross section properties for application with 6-Node triangles.

Figure 4.16 begins by defining the Gauss points and shape functions for a 6-Node

triangle.

Figure 4.16: Gauss Quadrature Weighting Functions [20, 21]

Figure 4.17: 6-Node triangle warping independent and dependent cross section properties

Figure 4.17 obtains from BCSP the y- and z- coordinates, as well as the warping

function for each node on the element from previous sections section of the code. With

this data, it is important calculate the Jacobian matrix, Je, and the shape function vector,

N, for each gauss point on the element (Figure 4.18). From here the finite element

formulations of Subsection 3.4.1.2 and 3.4.2.2 the warping independent and dependent

cross section properties are calculated.

83
Figure 4.18: 6-Node warping independent and dependent cross section properties (continued)

Figure 4.19: 6-Node Centroidal calculations and Parallel axis theorem

84
Figure 4.20: Unrestrained and Restrained Warping Cross Section Property Calculations

Figure 4.20 highlights the portion of the of the cross section property tool were

the warping independent and dependent properties are put to use in order to calculate

the torsion constant, J, shear center location, ys and zs, and the warping constant, Γ, if

the warping function approach has been taken. If the stress function approach is taken

however, only the torsional constant is obtained from this portion of the code.

4.3.2. Problem Conditioning: Stress Function vs. Warping Function

Since the BCSP code is designed for application to both the stress function and

the warping function approach to the problem of torsion, it is prudent to discuss how

external functions must change accordingly.

85
4.3.2.1. Stress Function

Given:

∇ 2ψ = −2Gθ (4.4)

With boundary conditions

ψ =0 (4.5)

on the external boundary.

For this problem, external function f.m must be set equal to 2 and external function

u_d.m must be set to zero. For the stress function approach only Dirichlet boundary

conditions are used to identify external boundaries.

As a result the expected outputs from BCSP for the solution of the torsion

problem with the stress function will be the warping independent cross section

properties and the torsion constant, J.

4.3.2.2. Warping Function

Given:

∇ 2ω = 0 (4.4)

with boundary conditions

ω=0 (4.5)

∂ω
=g (4.6)
∂n

The Dirichlet boundary condition of Equation 4.5 is necessary to mark areas of zero

warping on the cross section so that the stiffness matrix may be fixed in place [16] at a

point to provide a non-singular condition. The Neumann boundary condition must be

86
placed on all external and internal boundaries of the given cross section. For this

problem, external functions f.m and u_d.m must be set to zero.

As a result, the expected outputs from BCSP for the solution of the torsion

problem with the warping function, ω, will be the warping independent and warping

dependent cross section properties and the torsion constant, J, and the warping constant,

Γ.

87
CHAPTER 5

VALIDATION AND APPLICATION OF BCSP FEM TOOL

The development of a finite element tool for the calculation of beam cross

section properties has been presented from theoretical background to the finite element

formulation and implementation. The following chapter provides five example

applications of the BCSP FEM code; a circular cylinder, two square, multiply-

connected cross sections, a C-Beam, and an I-Beam. Following these brief applications

of the BCSP code, the circular cross section once again is presented, but this time to

compare the benefits of a 6-Node curved-sided triangular element as opposed to a 3-

Node straight-sided triangular element. A comparison of the stress function and the

warping function is also performed. Results from the validation problems are compared

to ANSYS in all cases and Reference [5] where available.

5.1 Circular Cylinder

The first application of the BCSP FEM code is toward a circular cylinder. As

mentioned previously, the circular cross section is the only cross section which does not

experience any warping, or out-of-plane displacement, under a torsional load.

Therefore, it is expected that the warping constant, Γ, will be zero.

Figure 5.1 provides the proper dimensions for the given problem, which is

solved using the warping function approach only, since the geometry is multiply

connected. For use in BCSP, this problem is modeled taking advantage of symmetry.

88
Problem Setup:

• Only the upper right quadrant is modeled.

• Dirichlet boundary conditions (ω = 0) are applied on the vertical and

horizontal lines of symmetry.

• Neumann boundary conditions have been applied to the internal and

external surfaces.

• Only 3-Node triangles are used.

• Warping function approach is used.

Figure 5.1: Circular Cylinder: 3-Node vs. 6-Node study

Table 5.1 contains results as obtained by ANSYS Beam Tool and BCSP for the

circular cylinder of Figure 5.1.

89
Table 5.1: Circular Cylinder Results: ANSYS and BCSP

A small discrepancy can be seen between ANSYS results and BCSP results.

The meshed cross section used to produce inputs files for BCSP was not fine enough,

however with only 338 elements still managed to remain accurate having an error of

only 1.41% on the torsional constant value. As expected, the warping constant is

verified by ANSYS and BCSP to be zero for a circular cross section.

5.2 Square Section with One Square Cut-Out

The next application is toward a square 2x2 cross section with a square 1x1 hole

cut out of its center as seen in Figure 5.2. Results from the BCSP code are compared in

Table 5.2 to those found using the ANSYS Beam Tool.

90
Figure 5.2: Square cross section with square hollow center

Problem Setup:

• Only the upper right quadrant is modeled.

• Dirichlet boundary conditions (ω = 0) are applied on the vertical and

horizontal lines of symmetry.

• Neumann boundary conditions have been applied to the internal and

external surfaces.

• Only 3-Node triangles are used.

• Warping function approach is used.

A mesh of 700 3-Node elements was placed on the cross section providing very

good agreement with the results obtained by ANSYS. There is only a 0.54% difference

between the values seen here for the warping constant, Γ, and 0.048% difference in the

torsional constant, J.

91
Table 5.2: Square cross section with square cut-out results

5.3 Square Section with Two Square Cut-Outs

The next application is toward a square 10x10 square cross section with two

square 3x8 spaces cut out. The two cut-out sections are centered symmetrically about

the centroid of the cross section with coordinates as given in Figure 5.3. Results from

the BCSP code are compared in Table 5.3 to those found using the ANSYS Beam Tool.

Problem Setup:

• Only the upper right quadrant is modeled.

• Dirichlet boundary conditions (ω = 0) are applied on the vertical and

horizontal lines of symmetry.

• Neumann boundary conditions have been applied to the internal and

external surfaces.

• Only 3-Node triangles are used.

• Warping function is approach used.

92
Figure 5.3: Square cross section with two hollow rectangular cut-outs

A mesh of 3,100 3-Node elements was placed on the cross section providing

very good agreement with the results obtained by ANSYS. There is only a 0.42%

difference between the values seen here for the warping constant, Γ, and 0.087%

difference in the torsional constant, J.

Table 5.3: Square cross section with two cut-outs results

93
5.4 C-Beam Application

The next application is toward a symmetric square C-Beam, an open section,

with height of 19 inches. and top and bottom flange length of 8.5 inches. See Figure 5.4

below. Results from the BCSP code are compared in Table 5.4 to those found using

the ANSYS Beam Tool and the Pilkey cross section tool, see Reference [5].

Problem Setup:

• Only the upper half of the cross section is modeled.

• Dirichlet boundary conditions (ω = 0) are applied on the horizontal line

of symmetry.

• Neumann boundary conditions have been applied to external surfaces.

• Only 3-Node triangles are used.

• Warping function approach is used.

Figure 5.4: Channel Beam

94
A mesh of 1,848 3-Node elements was placed on the cross section providing

very good agreement with the results obtained by ANSYS. There is almost exact

agreement between BCSP and ANSYS for this application, with the Pilkey code

differing only slightly. Any difference is so small that almost no effect is made.

Table 5.4: Channel Beam Results

5.5 I-Beam Application

The next application is toward a symmetric I-Beam, an open section, with height

of 20 inches. and top and bottom flange length of 10 inches. See Figure 5.5 below.

Results from the BCSP code are compared in Table 5.5 to those found using the

ANSYS Beam Tool.

• Only the upper right quadrant is modeled.

• Dirichlet boundary conditions (ω = 0) are applied on the vertical and

horizontal lines of symmetry.

• Neumann boundary conditions have been applied to external surfaces.

• Only 3-Node triangles are used.

• Warping function is approach used.

95
Figure 5.5: I-Beam

A mesh of 2,278 3-Node elements was placed on the cross section providing

very good agreement with the results obtained by ANSYS. There is a 0.04% difference

between the values seen here for the warping constant, Γ, and 0.002% difference in the

torsional constant, J.

Table 5.5: I-Beam results

96
5.6 Circular Cylinder Application: 3-Node vs. 6-Node Triangles

The same circular cylinder discussed in Section 5.1 is repeated here in Figure

5.7 to be solved using first 3-Node straight-edged triangular elements and then 6-Node

curved-sided elements. Application of straight edged elements to curved boundaries

requires the user to generate a very fine mesh of the cross section to better approximate

curvature. The advantage of the 6-Node curved-sided element is that it can easily form

to a curved shape allowing for a better approximation with much fewer elements. See

Figure 5.6.

Figure 5.6: 3-Node straight-sided and 6-Node curved-sided elements

Figure 5.7: Circular Cylinder

97
The circular cylinder was fit with a mesh of 136 triangular straight-sided

elements in one case and only 5 6-Node curved-sided elements in another. In spite of

the low number of 6-Node elements used, the results obtained are as good as or better

than the 3-Node triangles because the element could form to the curved surface. See

Table 5.6.

Table 5.6: Circular Cylinder: 3-Node vs. 6-Node results

From the table above it can be seen that there is only a 0.05% difference

between BCSP with the 6-Node element and ANSYS in the value of the torsional

constant. The 3-Node element however has a slightly less accurate answer in

comparison to ANSYS having a difference of 1% between the torsional constant, J.

5.7 Stress Function vs. Warping Function

Finally, it is of interest to contrast the application of the stress function approach

to the warping function approach for the calculation of the torsional constant. For the

rectangular cross section in Figure 5.8, both the stress function and warping function

were employed to solve for the torsional constant using a varying amount of 3-Node

triangular elements.

98
Figure 5.8: Rectangular C-S: Stress Function vs. Warping Function

Figure 5.9 shows the convergence of both cases toward the real value of the

torsional constant as calculated in ANSYS. The green curve represents the warping

function and is seen to converge quickly in comparison to the stress function approach;

red curve. After approximately 1000 elements the warping function converges to

within 0.054% of the value given by ANSYS. The stress function takes over 4000

elements to arrive at an appropriate value of the torsional constant. The reason for the

warping function’s quicker convergence is because it is a method based in displacement

while the stress function is based in a force formulation [5].

99
Figure 5.9: Convergence of Stress Function vs. Warping Function on Torsional Constant

Results of the stress function and the warping function solution compared to ANSYS

are available in Figure 5.7.

Table 5.7: Rectangular C-S: Stress Function vs. Warping Function

100
CHAPTER 6

CONCLUSIONS AND RECOMMENDATIONS

The purpose of this project has been accomplished through the creation of a

finite element tool that calculates fundamental warping independent and warping

dependent cross section properties for the case of pure torsion. The tool has been

created in MATLAB and is able to handle open or closed cross sections for which at

least one point of zero warping is known. Though the ability to determine the cross

section properties of a cross section with an arbitrary shape was desired, the restrictions

placed on the problem at hand by the stiffness matrix prevented this. To prevent the

stiffness matrix from being singular, at least one point of zero warping is needed in

order to apply the Dirichlet boundary condition. By securing the stiffness matrix in this

manner, the solution should be non-singular, thereby allowing a solution to be

determined.

The code has been tested and validated through the use of five cross sections for

which the properties determined by the code were compared to the same results

determined from ANSYS. The results match extremely well, differing by less than one

percent with a fine mesh. Also, a study has been conducted which compared the results

obtained using the code with 3-Node straight-edge triangles and with 6-Node curve-

sided triangles to the results obtained from ANSYS. The 6-Node triangles are able to fit

a curved surface, therefore, for the same level of accuracy, the 6-Node triangle mesh

101
required less elements (5) than the 3-Node triangles (136) to reach the result. Finally, a

study has been completed which compared the results obtained with the stress function

versus the results obtained with the warping function. The results show that the stress

function requires significantly more elements than the warping function to arrive at the

same level of accuracy in comparison to the results returned by ANSYS, 4000 versus

1000 elements, respectively.

In conclusion, it has been determined that the warping function is the best

approach for a project of this type, as it allows for the calculation of warping dependent

cross section properties that include the warping constant as well as the torsional

constant. In contrast, the stress function only allows for the calculation of the torsional

constant. Additionally, the case studies of this project show that 6-Node, curve-sided

triangular element is more efficient for applications with curved surfaces than the 3-

Node straight-edge triangles.

At the conclusion of this project, there are a few recommendations for further

progress in the calculation of cross section properties. First, it should be investigated

how the stiffness matrix can be non-singular naturally as Reference [5] suggests; that is,

how the stiffness matrix can be used to solve the warping function without the

restriction of a known location of zero warping for Dirichlet boundary conditions. With

this ability, a more in-depth comparison of the results returned through the warping

function and stress function can be conducted.

102
Secondly, it would be beneficial to modify the code to allow input for an

applied moment or torque, as it would allow for further calculation of the angle of twist,

the shear stresses, and the warping normal stress of a cross section.

Lastly, the theory of transverse shear should be incorporated into the code to further

calculate cross section properties of shear deformation coefficients and the shear center

without using Trefftz’s shear center approximation.

103
APPENDIX A

BCSP FINITE ELEMENT TOOL: MATLAB SOURCE CODE

104
%% BCSP applies the finite element method to Laplace or Poisson's
Equation.
%
%
clear
%
% Read and load the nodal coordinate input file.
%
load coordinates.dat;
%
% Read and load the 3-Node and 6-Node triangle input files.
%
load elements3.dat;
load elements6.dat;
%
% Read and load the Neumann boundary condition input file.
% EVAL command is to create an empty NEUMANN array if no neumann file
% is found.
%
eval ( 'load neumann.dat;', 'neumann=[];' );
%
% Read and load the Dirichlet boundary condition input file.
%
load dirichlet.dat;

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%
% Set problem symmetry type:
% =0: Zero Symmetry
% =1: Symmetric about Y-Axis
% =2: Symmetric about Z-Axis
% =4: Symmetric about Y and Z-Axis
% Ex: If value=4 then the problem can be modeled as a single
Quadrant
% with final answers being adjusted accordingly.
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%

sym=input('Set symmetry value: [No sym->0; sym y->1; sym z->2; y & z-
>4] ');

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%
% Stiffness Matrix Assembly
%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%

A = sparse ( size(coordinates,1), size(coordinates,1) );

105
% 3-Node Triangles
for j = 1 : size(elements3,1)
A(elements3(j,:),elements3(j,:)) =
A(elements3(j,:),elements3(j,:)) ...
+ stima3(coordinates(elements3(j,:),:));
end

% 6-Node Triangles
for j = 1 : size(elements6,1)

A(elements6(j,:),elements6(j,:)) =
A(elements6(j,:),elements6(j,:)) ...
+ stima6(coordinates(elements6(j,:),:));

end

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%
% Assemble Volume Forces
%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%

b = sparse ( size(coordinates,1), 1 );

% Three Node Trianlge


for j = 1 : size(elements3,1)
b(elements3(j,:)) = b(elements3(j,:)) ...
+ det( [1,1,1; coordinates(elements3(j,:),:)'] ) * ...
f(sum(coordinates(elements3(j,:),:))/3)/6;
end

%Six Node Triangle


for j = 1 : size(elements6,1)
b(elements6(j,:)) = b(elements6(j,:)) ...
+ volume6(coordinates(elements6(j,:),:))* ...
f(sum(coordinates(elements6(j,:),:))/4)/4;
end

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%
% Neumann Boundary Conditions
%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%

if ( ~isempty(neumann) ) % Checks to see if neumann.dat is empty

for j=1:size(neumann,1)

106
% Application of normal warping derivative 'g' toward the Right Hand
Side
% through the addition of (E/2)*g where E is the length of the side
% represented by each neumann.dat node pair

b(neumann(j,:)) = b(neumann(j,:)) + ...


norm(coordinates(neumann(j,1),:) -
coordinates(neumann(j,2),:)) * ...
warp(coordinates(neumann(j,:),:))/2;
end
end

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%
% Apply Dirichlet Boundary Conditions
%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%
%
% Determine which nodes are associated with Dirichlet conditions.
% Assign the corresponding entries of u, and adjust the right hand
side.
%
u = sparse ( size(coordinates,1), 1 );
BoundNodes = unique ( dirichlet );
u(BoundNodes) = u_d ( coordinates(BoundNodes,:) );
b = b - A * u;

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%
% Solve A * u = b for the remaining unknown values of u.
%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%

FreeNodes = setdiff ( 1:size(coordinates,1), BoundNodes );

u(FreeNodes) = A(FreeNodes,FreeNodes) \ b(FreeNodes);

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%
% Cross Section Properties
%

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%
u; % u= nodal warping function value if f=0

107
% u= nodal stress function value if f=2

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%
% Warping Independent and Dependent Cross Section Properties
%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%

Area=0; Qyi=0; Qzi=0; Qwi=0; Iyi=0; Izi=0; Iyzi=0; Iwi=0; Iywi=0;


Izwi=0;
Jstress=0;

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%
%%%%%%%%%%%%%%%%%%%% 3-Node Triangle Calculations
%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%

for j=1:size(elements3,1)

y1=coordinates(elements3(j,1),1);
y2=coordinates(elements3(j,2),1);
y3=coordinates(elements3(j,3),1);
z1=coordinates(elements3(j,1),2);
z2=coordinates(elements3(j,2),2);
z3=coordinates(elements3(j,3),2);

u1=u(elements3(j,1));
u2=u(elements3(j,2));
u3=u(elements3(j,3));

Ae=0.5*((y2*z3-y3*z2)+(y1*(z2-z3))+(z1*(y3-y2)));
Area=Area+Ae;

Qyi=Qyi+((Ae/3)*(z1+z2+z3));
Qzi=Qzi+((Ae/3)*(y1+y2+y3));
Qwi=Qwi+((Ae/3)*(u1+ u2 + u3));

Iyi=Iyi+((Ae/6)*((z1^2)+(z2^2)+(z3^2)+(z1*z2)+(z2*z3)+(z1*z3)));
Izi=Izi+((Ae/6)*((y1^2)+(y2^2)+(y3^2)+(y1*y2)+(y2*y3)+(y1*y3)));

Iyzi=Iyzi+((Ae/6)*((y1*z1)+(y2*z2)+(y3*z3)))+((Ae/12)*((y1*z2)+(y2*z1)
...
+(y1*z3)+(y3*z1)+(y2*z3)+(y3*z2)));

Iwi=Iwi + ((Ae/6)*((u1^2)+(u2^2)+(u3^2)+(u1*u2)+(u2*u3)+(u1*u3)));

108
Iywi=Iywi +
((Ae/6)*((y1*u1)+(y2*u2)+(y3*u3)))+((Ae/12)*((y1*u2)+(y2*u1)...
+(y1*u3)+(y3*u1)+(y2*u3)+(y3*u2)));
Izwi=Izwi +
((Ae/6)*((z1*u1)+(z2*u2)+(z3*u3)))+((Ae/12)*((z1*u2)+(z2*u1)...
+(z1*u3)+(z3*u1)+(z2*u3)+(z3*u2)));

if (f == 2)
Jstress = Jstress + (2/3)*Ae*sum(u(elements3(j,:)));% Torsion
Constant for stress function approach
end

end

%The subcript i denotes 'initial' values before symmetry is applied

Area; Qyi; Qzi; Qwi; Iyi; Izi; Iyzi; Iwi; Iywi; Izwi;

Yci=Qzi/Area;
Zci=Qyi/Area;

Iyci=Iyi-(Area*(Zci^2)); %Parallel Axis Theorem


Izci=Izi-(Area*(Yci^2)); %Parallel Axis Theorem
Iyzci=Iyzi-(Area*Yci*Zci); %Parallel Axis Theorem

Iywci=Iywi-(Area*Yci*Zci); %Parallel Axis Theorem


Izwci=Izwi-(Area*Yci*Zci); %Parallel Axis Theorem

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%
%%%%%%%%%%%%%%%%%%%% 6-Node Triangle Calculations
%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%

%W=Vector containing Gauss Point values for each node


%Z=Vector containing isoparametric shape function value for each node
%E=Vector containing isoparametric shape function value for each node

W=[0.109951743655322, 0.109951743655322, 0.109951743655322,


0.223381589678011, 0.223381589678011, 0.223381589678011];
Z=[0.091576213509771, 0.816847572980459, 0.091576213509771,
0.445948490915965, 0.108103018168070, 0.445948490915965];
E=[0.091576213509771, 0.091576213509771, 0.816847572980459,
0.445948490915965, 0.445948490915965, 0.108103018168070];

for j=1:size(elements6,1)

109
y(1,1)=coordinates(elements6(j,1),1);
y(2,1)=coordinates(elements6(j,2),1);
y(3,1)=coordinates(elements6(j,3),1);
y(4,1)=coordinates(elements6(j,4),1);
y(5,1)=coordinates(elements6(j,5),1);
y(6,1)=coordinates(elements6(j,6),1);
z(1,1)=coordinates(elements6(j,1),2);
z(2,1)=coordinates(elements6(j,2),2);
z(3,1)=coordinates(elements6(j,3),2);
z(4,1)=coordinates(elements6(j,4),2);
z(5,1)=coordinates(elements6(j,5),2);
z(6,1)=coordinates(elements6(j,6),2);

u1(1,1)=u(elements6(j,1)); u1(2,1)=u(elements6(j,2));
u1(3,1)=u(elements6(j,3)); u1(4,1)=u(elements6(j,4));
u1(5,1)=u(elements6(j,5)); u1(6,1)=u(elements6(j,6));

temp=0;temp1=0;temp2=0;temp3=0;temp4=0; temp5=0; temp6=0; temp7=0;


temp8=0; temp9=0;
for i=1:6
N(1,1)=1+(2*((Z(1,i))^2))+(2*((E(1,i))^2))+(4*Z(1,i)*E(1,i))-
(3*Z(1,i))-(3*E(1,i));
N(1,2)=(2*((Z(1,i))^2)-Z(1,i));
N(1,3)=(2*((E(1,i))^2)-E(1,i));
N(1,4)=(4*Z(1,i)-4*((Z(1,i))^2)-4*Z(1,i)*E(1,i));
N(1,5)=4*Z(1,i)*E(1,i);
N(1,6)=(4*E(1,i)-4*((E(1,i))^2)-4*Z(1,i)*E(1,i));

J(1,1)= (-3+4*Z(1,i)+4*E(1,i))*y(1,1)+(4*Z(1,i)-
1)*y(2,1)+0*y(3,1)+...
(4-8*Z(1,i)-4*E(1,i))*y(4,1)+4*E(1,i)*y(5,1)+(-
4*E(1,i))*y(6,1);
J(1,2)= (-3+4*Z(1,i)+4*E(1,i))*z(1,1)+(4*Z(1,i)-
1)*z(2,1)+0*z(3,1)+...
(4-8*Z(1,i)-4*E(1,i))*z(4,1)+4*E(1,i)*z(5,1)+(-
4*E(1,i))*z(6,1);
J(2,1)= ((-3+4*Z(1,i)+4*E(1,i))*y(1,1)+ 0*y(2,1)+(4*E(1,i)-
1)*y(3,1)+...
(-4*Z(1,i))*y(4,1)+(4*Z(1,i))*y(5,1)+(4-4*Z(1,i)-
8*E(1,i))*y(6,1));
J(2,2)= ((-3+4*Z(1,i)+4*E(1,i))*z(1,1)+ 0*z(2,1)+(4*E(1,i)-
1)*z(3,1)+...
(-4*Z(1,i))*z(4,1)+(4*Z(1,i))*z(5,1)+(4-4*Z(1,i)-
8*E(1,i))*z(6,1));
Jdet=det(J);

Ae = temp + 0.5*(Jdet*W(1,i));
Qye = temp1 + 0.5*(Jdet*W(1,i)*N*z);
Qze = temp2 + 0.5*(Jdet*W(1,i)*N*y);
Iye = temp3 + 0.5*(Jdet*W(1,i)*((N*z)^2));

110
Ize = temp4 + 0.5*(Jdet*W(1,i)*((N*y)^2));
Iyze = temp5 + 0.5*(Jdet*W(1,i)*((N*y)*(N*z)));
Qwe = temp6 + 0.5*(Jdet*W(1,i)*(N*u1));

Iwe = temp7 + 0.5*(Jdet*W(1,i)*((N*u1)^2));


Iywe = temp8 + 0.5*(Jdet*W(1,i)*((N*y)*(N*u1)));
Izwe = temp9 + 0.5*(Jdet*W(1,i)*((N*z)*(N*u1)));

temp=Ae; temp1=Qye; temp2=Qze; temp3=Iye; temp4=Ize;


temp5=Iyze; temp6=Qwe;
temp7=Iwe; temp8=Iywe; temp9=Izwe;
end
Area= Area + Ae;
Qyi= Qyi + Qye;
Qzi= Qzi + Qze;
Iyi= Iyi + Iye;
Izi= Izi + Ize;
Iyzi= Iyzi + Iyze;
Qwi= Qwi + Qwe;
Iwi= Iwi + Iwe;
Iywi= Iywi + Iywe;
Izwi= Izwi + Izwe;
end

Area;
Qyi;
Qzi;
Qwi;
Iywi;
Izwi;
Yci=Qzi/Area;
Zci=Qyi/Area;
Iyi;
Izi;
Iyzi;
Qwi;

% Calculate global values for the full geometry based on given


symmetry

if (sym == 0) % No Symmetry
symmetry=sym+1
Qy=Qyi, Qz=Qzi, Qw=Qwi
Yc=Yci, Zc=Zci
Iyc=Iyi-(Area*(Zci^2)), Izc=Izi-(Area*(Yci^2))
Iyzc=Iyzi-(Area*Yci*Zci)
Areatotal=Area

elseif (sym == 1 ) % Symmetric about Y-Axis


symmetry=sym+1;
Areatotal=symmetry*Area

111
Qy=Qyi-Qyi, Qz=Qzi*symmetry;
Yc=Qz/Areatotal
Zc=Qy/Areatotal
Qw=Qwi-Qwi;
Iy=Iyi*symmetry;
Iz=Izi*symmetry;
Iyz=Iyzi-Iyzi;
Iyc=Iy-(Areatotal*(Zc^2))
Izc=Iz-(Areatotal*(Yc^2))
Iyzc=Iyz-Iyz
Iw=Iwi*symmetry
Iyw=Iywi-Iywi
Izw=Izwi*symmetry

elseif (sym == 2) % Symetric about Z-Axis


symmetry=sym
Yc=Yci-Yci
Zc=Zci
Qw=Qwi-Qwi
Iy=Iyi*symmetry;
Iz=Izi*symmetry;
Iyz=Iyzi-Iyzi;
Iyc=Iy-(Areatotal*(Zc^2))
Izc=Iz-(Areatotal*(Yc^2))
Iyzc=Iyzi-Iyzi
Iw=Iwi*symmetry
Iyw=Iywi*symmetry
Izw=Izwi-Izwi
Areatotal=symmetry*Area

elseif (sym ==4) % Symetric about Y and Z-Axis


symmetry=sym;
Areatotal=symmetry*Area
Yc=Yci-Yci+Yci-Yci
Zc=Zci-Zci+Zci-Zci
Qw=Qwi-Qwi+Qwi-Qwi
Iy=Iyi*symmetry
Iz=Izi*symmetry
Iyc=Iy-(Areatotal*(Zc^2))
Izc=Iz-(Areatotal*(Yc^2))
Iyzc=Iyzi-Iyzi+Iyzi-Iyzi
Iw=Iwi*symmetry
Iyw=Iywi-Iywi+Iywi-Iywi
Izw=Izwi-Izwi+Izwi-Izwi
end

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%
% Unrestrained and Restrained Warping Cross Section Properties
%

112
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%

if (f == 0) %Warping Function Approach

J=Iy+Iz-symmetry*(b'*u) % Torsion Constant

Ys=((Iyzc*Iyw)-(Izc*Izw))/((Iyc*Izc)-(Iyzc^2)) %Shear Center


Y-Coord
Zs=((Iyc*Iyw)-(Iyzc*Izw))/((Iyc*Izc)-(Iyzc^2)) % Shear Center
Z-Coord
Gamma= Iw-((Qw^2)/Areatotal)-Zs*Iyw+Ys*Izw %Warping Constant

ustar = sparse ( size(coordinates,1), 1 );


if ( ~isempty(elements3) )
warp=unique(elements3);
end
if ( ~isempty(elements6) )
warp=unique(elements6);
end
%Restrained Warping Function
ustar(warp)= u-
(Zs*(coordinates(warp,1)))+(Ys*(coordinates(warp,1)))-(Qw/Areatotal);

end
if (f == 2)
Jstress=Jstress*symmetry
end

113
REFERENCES

[1] Little, Robert W. Elasticity. Prentice-Hall, 1973.


[2] Archon Engineering Software. <www.archoneng.com>.
[3] Olver, Peter J. Applied Mathematics Lecture Notes. 2007. Chapter 15: the Planar
Laplace Equation.
[4] Reddy, J N. An Introduction to the Finite Element Method. 3rd ed. McGraw-Hill,
2006.
[5] Pilkey, Walter D. Analysis and Design of Elastic Beams: Computational Mehtods.
John Wiley& Sons, 2002.
[6] ANSYS User Guide. ANSYS.
[7] Gere, James M. Mechanics of Materials. Brooks/Cole, 2001.
[8] Timoshenko, S P., and J N. Goodier. Theory of Elasticity. 3rd ed. McGraw-Hill,
1970.
[9] Cook, Robert D., David S. Malkus, Michael E. Plesha, and Robert J. Witt. Concepts
and Applications of Finite Element Analysis. 4th ed. John Wiley & Sons, 2002.
[10] Atkinson, Kendall E. Theoretical Numerical Analysis: a Functional Analysis
Framework. Springer, 2001.
[11] Alberty, Jochen, Carsten Carstensen, and Stefan A. Funken. "Remarks Around 50
Lines of Matlab: Short Finite Element Implementation." Numerical Algorithms
20 (1999): 117-137.
[12] Ciarlet, Philippe G. The Finite Element Method for Elliptic Problems. North-
Holland Company, 1978.
[13] Braess, Dietrich. Finite Elements: Theory, Fast Solvers, and Applications in Solid
Mechanics. Cambridge UP, 2007.
[14] Segerlind, Larry J. Applied Finite Element Analysis. John Wiley & Sons, 1984.
[15] Lawrence, Kent. "Element Formulation." ME5310/AE5330 - Finite Element
Methods. University of Texas At Arlington.
<http://mae.uta.edu/~lawrence/me5310/me5310.htm>.
[16] Lauterbach, Gerald F., and Joeseph L. Krahula. "A Finite Element Solution for
Saint-Venant Torsion." AIAA JOURNAL 7 (1969): 2200-2203.
[17] Mathworks. MatLab and Simulink. Natick: The Mathworks Inc., 2007.
[18] Kattan, Peter Issa. MATLAB Guide to Finite Elements : An Interactive Approach.
New York: Springer, 2002.
[19]"EasyMesh version 1.4." EasyMesh. <http://www-
dinma.univ.trieste.it/nirftc/research/easymesh/easymesh.html>.
[20] Upendrakumar Soni, Sachin. "FEM Element Formulation Efficiency." Master of
Engineering in Mechanical Engineering Report, The University of Texas at
Arlington, Department of Mechanical and Aerospace Engineering, Nov 2005.

114
[21] Burnett, David S. Finite Element Analysis : From Concepts to Applications. New
York: Addison-Wesley Longman, Limited, 1987.

115
BIOGRAPHICAL INFORMATION

Bryan Mixon was born in Oklahoma in 1981. Bryan graduated from Mustang

High School in 2000 and went on to pursue an undergraduate degree at the University

of Oklahoma. In 2005, Bryan graduated with distinction, earning a Bachelor of Science

in Aerospace Engineering from the University of Oklahoma. Continuing with his

education, Bryan moved to Texas in pursuit of a Master of Science in Aerospace

Engineering at the University of Texas at Arlington. While at UTA, Bryan was a

member of the Phi Kappa Phi National Honor Society, and attended two AIAA Annual

Aerospace Sciences Meeting and Exhibit and Author and Co-Authored three conference

publications from 2006 to 2007.

116

Anda mungkin juga menyukai