Anda di halaman 1dari 192

AN INTEGRATED METHODOLOGY FOR

STATISTICAL TOLERANCE ANALYSIS


OF FLEXIBLE ASSEMBLIES

by
Alan J. Mortensen

A thesis submitted to the faculty of


Brigham Young University
in partial fulfillment of the requirements for the degree of
Master of Science

Department of Mechanical Engineering


Brigham Young University
August 2002

ii

ABSTRACT
A METHODOLOGY FOR STATISTICAL TOLERANCE ANALYSIS
OF FLEXIBLE ASSEMBLIES
Alan J. Mortensen
Department of Mechanical Engineering
Master of Science
Statistical tolerance analysis has become a vital link between product design and
manufacture, aiding in the constant drive to increase quality, while lowering production
costs. Although conventional tolerance analysis methods are limited by the assumption
of part rigidity, recent steps have been taken to broaden this field of study to include
flexible parts. Such steps include the combination of statistical tolerance analysis with
finite element analysis to predict closure forces, deformations and internal stresses
resulting from assembly processes. The achievement of such would greatly aid
aerospace, automotive and other industries that routinely use compliant parts in design.
The objective of this thesis is to present an integrated methodology for the statistical
tolerance analysis of flexible assemblies. To accomplish this objective, four main tasks
were completed:
1. Integration of rigid STA with FEA for variation analysis of assemblies with
compliant parts.
2. Use of commercial FEA software for modeling of compliant assemblies.
3. Statistical FEA, incorporating both material (elastic coupled) and geometric
(continuity coupled) covariance.
4. Inclusion of both compliant and rigid covariance components.
The result of this research is an integrated 4-step process that allows for flexible assembly
statistical tolerance analysis (FASTA). The FASTA process is presented in detail and
summarized graphically using process flow diagrams. Groundwork is also presented for
the treatment of form variation at assembly joints containing flexible parts. It is hoped
that the methods presented will become a valuable tool for analysis in the design of
flexible assemblies.

iii

iv

ACKNOWLEDGEMENTS
Special thanks must be given to my wonderful wife who has supported me and sacrificed
to help me to accomplish this work. I would also like to thank Dr. Chase for his
mentoring and guidance in this undertaking.

vi

Table of Contents:
CHAPTER 1

................................................................................................................. 1

1.1

Introduction......................................................................................................... 1

1.2

Overview of New Methods................................................................................. 3

1.3

Thesis Objectives ................................................................................................ 8

1.4

Procedure ............................................................................................................ 8

1.5

Scope Delimitations ............................................................................................ 9

1.5.1

Linear analysis ............................................................................................ 9

1.5.2

Statistical assumptions ................................................................................ 9

1.5.3

FEA modeling........................................................................................... 10

1.6

Summary ........................................................................................................... 10

CHAPTER 2
2.1

............................................................................................................... 11

Background ....................................................................................................... 11

2.1.1

Tolerance analysis of rigid assemblies ..................................................... 11

2.1.2

Tolerance analysis of flexible parts .......................................................... 14

2.2

Summary ........................................................................................................... 18

CHAPTER 3

............................................................................................................... 19

3.1

Introduction....................................................................................................... 19

3.2

Sources of Assembly Variation ........................................................................ 19

3.3

Vector Loop Modeling...................................................................................... 21

3.3.1

DLM in 3D with dimensional and kinematic variation ............................ 21


vii

3.3.1.1

Closed vector loops............................................................................... 24

3.3.1.2

Open vector loops ................................................................................. 26

3.3.2

Including form variation in 3D DLM ....................................................... 27

3.3.2.1

Form variation at Planar and Cylindrical Slider joints ......................... 27

3.3.2.2

Dimensional, kinematic, & form variation in DLM ............................. 30

3.3.2.3

Sensitivities and kinematic variation estimates .................................... 31

System............................................................................................................... 32
Closed Loop ...................................................................................................... 32
Open Loop ........................................................................................................ 32
3.3.3
3.4

Vector loop modeling practices ................................................................ 32

Summary ........................................................................................................... 33

CHAPTER 4

............................................................................................................... 35

4.1

Introduction....................................................................................................... 35

4.2

Modeling Flexible Assemblies Using DLM ..................................................... 37

4.2.1

Obtaining the mean gap solution .............................................................. 37

4.2.2

Modeling flexible gap variation................................................................ 38

4.2.3

Resolving gap variation into components................................................. 42

4.2.4

Example assembly .................................................................................... 51

4.3

4.2.4.1

Vector loop model of example assembly.............................................. 53

4.2.4.2

DLM applied to example assembly ...................................................... 54

4.2.4.3

Rotational and surface variation in example assembly......................... 60

Other Considerations for Flexible Assemblies ................................................. 62

4.3.1

Positional Variation of Nonparallel surfaces ............................................ 62


viii

4.3.2

Tooling contribution ................................................................................. 64

4.3.3

Multiple gap variation propagation........................................................... 65

4.3.4

Obtaining geometry and variation data..................................................... 67

4.4

Summary ........................................................................................................... 68

CHAPTER 5

............................................................................................................... 71

5.1

Introduction....................................................................................................... 71

5.2

Analysis Assumptions....................................................................................... 72

5.3

Background ....................................................................................................... 73

5.3.1

Matrix reduction........................................................................................ 75

5.3.2

Equivalent stiffness................................................................................... 77

5.4

FEA Modeling of Flexible Parts ....................................................................... 79

5.4.1

Define nominal geometry ......................................................................... 80

5.4.2

Material properties .................................................................................... 81

5.4.3

Element selection ...................................................................................... 82

5.4.3.1

Shell elements ....................................................................................... 83

5.4.4

Mesh part .................................................................................................. 84

5.4.5

Reduced stiffness matrix output................................................................ 85

5.5

Summary ........................................................................................................... 88

CHAPTER 6

............................................................................................................... 91

6.1

Introduction....................................................................................................... 91

6.2

Step 3: Covariance Calculation......................................................................... 92

6.2.1

Introduction to covariance ........................................................................ 93

6.2.2

Surface continuity models......................................................................... 95


ix

6.2.2.1

Sinusoidal model................................................................................... 96

6.2.2.2

Polynomial series models ..................................................................... 98

6.2.2.3

Hybrid method .................................................................................... 101

6.2.3

Rigid contribution to gap covariance...................................................... 103

6.2.4

Gap covariance of the example assembly............................................... 106

6.2.5

Step 3 process summary.......................................................................... 110

6.3

Step 4: Statistical FEA Solution ..................................................................... 113

6.3.1

Statistical solution of assembly force ..................................................... 114

6.3.2

Mean solution of internal deformation and stress................................... 116

6.3.2.1

Mean deformation solution ................................................................. 117

6.3.2.2

Mean stress solution............................................................................ 118

6.3.3
6.3.3.1

Covariant deformation solution .......................................................... 120

6.3.3.2

Covariant stress solution ..................................................................... 122

6.3.4
6.4

Covariance equations for deformation and stress ................................... 120

Step 4 process summary.......................................................................... 123

Chapter Summary ........................................................................................... 126

CHAPTER 7

............................................................................................................. 127

7.1

Introduction..................................................................................................... 127

7.2

Review of Rigid Form Variation Modeling.................................................... 127

7.3

Flexible Parallel to Rigid Form Variation Model ........................................... 129

7.4

Multiple Joint FASTA for Form Variation Inclusion ..................................... 131

7.5

Modes of Variation Propagation..................................................................... 133

7.6

Summary ......................................................................................................... 135


x

CHAPTER 8

............................................................................................................. 137

8.1

Introduction..................................................................................................... 137

8.2

Presentation of Assembly ............................................................................... 137

8.3

Setup of FASTA ............................................................................................. 139

8.3.1

Step 1: Determine misalignment............................................................ 139

8.3.2

Step 2: Model compliant parts ............................................................... 143

8.3.3

Step 3: Covariance calculation............................................................... 145

8.3.4

Step 4: Statistical FEA solution ............................................................. 146

8.4

Measurement for Analysis & Verification...................................................... 147

8.4.1

Measurement to determine fastening locations....................................... 147

8.4.2

Measurement for analysis verification.................................................... 149

8.5

Summary ......................................................................................................... 149

CHAPTER 9

............................................................................................................. 151

9.1

Introduction..................................................................................................... 151

9.2

Research Objectives........................................................................................ 151

9.3

Thesis Contributions ....................................................................................... 154

9.4

Suggested Future Work................................................................................... 155

9.5

Closing Statement ........................................................................................... 157

APPENDIX..159
APPENDIX A: Vector Loop Modeling Rules and Regulations for 2-D....161
APPENDIX B: FEA Model Issues and Suggestions..163
APPENDIX C: Sub-structuring in ANSYS for Stiffness Matrix Output.....169
REFERENCES173
xi

Table of Figures:
Figure 1-1: Illustration of flexible assembly deformation .................................................. 2
Figure 1-2 Process flow diagram of FASTA process ........Error! Bookmark not defined.
Figure 3-1 Kinematic Assembly Variation [Chase 1996]................................................ 20
Figure 3-2 Simple Vector Loop Model............................................................................. 21
Figure 3-3 3D Effects of form variation [Chase 1996].................................................... 28
Figure 3-4 Degrees of freedom for kinematic motions and geometric feature variations
[Chase 1996] ............................................................................................................. 29
Figure 4-1 Step 1 process summary.................................................................................. 36
Figure 4-2 Modeling multiple fastening points using variable vectors (fi, gi).................. 39
Figure 4-3 Form variation at multiple fastening points .................................................... 40
Figure 4-4 Example of modes of variation contribution................................................... 43
Figure 4-5 Dimensioned flexible part used for illustration............................................... 44
Figure 4-6 Illustration of translational variation in flexible part ...................................... 44
Figure 4-7 Illustration of rotational variation in flexible part........................................... 46
Figure 4-8 Form tolerance limits of surface variation ...................................................... 48
Figure 4-9 Example flexible assembly ............................................................................ 52
Figure 4-10 Example part dimensions ............................................................................. 53
Figure 4-11 Vector loop model of example assembly ...................................................... 54
Figure 4-12 Nonparallel surface variation ........................................................................ 62
Figure 4-13 Deformation of cantilever beams [Liu 1995]................................................ 65
xii

Figure 4-14 Order of assembly example - Trampoline..................................................... 66


Figure 4-15 Equilibrium trampoline assembly ................................................................. 66
Figure 4-16 Process flow diagram for statistically determining misalignment ................ 70
Figure 5-1 Step 2 process summary.................................................................................. 72
Figure 5-2 Example of stiffness of bar in tension............................................................. 73
Figure 5-3 Example of stiffness for 2-D bending ............................................................. 74
Figure 5-4 Boundary and interior nodes of a plate in bending [Tonks 2002]................... 76
Figure 5-5 Spring example of equivalent stiffness [Merkley 1998] ................................. 78
Figure 5-6 Nominal model of flexible parts in example assembly (created in ANSYS) 81
Figure 5-7 Examples of 3-D structural elements in ANSYS .......................................... 82
Figure 5-8 Suggested Shell Elements for flexible part modeling ..................................... 83
Figure 5-9 Flexible parts in example assembly map meshed with a 0.5 inch
edge-length................................................................................................................ 85
Figure 5-10 Master DOF and constraints applied to parts A and B.................................. 86
Figure 5-11 Reduced stiffness matrices for parts A and B (scale in pounds/inch)........... 87
Figure 5-12 Equivalent stiffness matrix (scale in pounds/inch) ....................................... 88
Figure 5-13 Process flow diagram for FEM of compliant parts ....................................... 90
Figure 6-1 Step 3 process summary.................................................................................. 93
Figure 6-2 Example of correlation between variables x and y ......................................... 94
Figure 6-3 Sinusoidal curve fit of independent points [Tonks 2002] ............................... 96
Figure 6-4 Orthogonal polynomial plots [Tonks 2002].................................................... 99
Figure 6-5 Rigid translational variation.......................................................................... 104
Figure 6-6 Rigid rotational variation .............................................................................. 105
Figure 6-7 Geometric covariance solutions for example assembly gap surfaces ........... 108
xiii

Figure 6-8 Rigid gap covariance solutions for example assembly ................................. 108
Figure 6-9 Gap covariance matrix solution and contributing components, for example
assembly.................................................................................................................. 109
Figure 6-10 Process flow diagram for calculating gap covariance................................. 112
Figure 6-11 Step 4 process summary.............................................................................. 113
Figure 6-12 Force covariance matrix for example assembly.......................................... 115
Figure 6-13 Plots of mean deformation and stress for example parts in ANSYS ........ 119
Figure 6-14 Displacement variance surface plot of the horizontal surface of
example part A........................................................................................................ 122
Figure 6-15 Process flow diagram for statistical FEA soltion........................................ 125
Figure 7-1 Example of rigid form variation propagation................................................ 128
Figure 7-2 Comparison of rigid and flexible joints ........................................................ 130
Figure 7-3 Process flow diagram of multiple joint FASTA application......................... 133
Figure 7-4 Variation modes at flexible joints for example assembly ............................. 134
Figure 8-1 Pictures of parts and fixture for leading edge assembly ............................... 138
Figure 8-2 Leading edge assembly gap at fastening nodes............................................. 140
Figure 8-3 Generic vector loop describing the gap distance at node i............................ 140
Figure 8-4 Preliminary models of leading edge flexible parts, made in ANSYS ......... 145
Figure 8-5 Assembly fixture positioned on CMM.......................................................... 148
Figure A-1 Vector loop joint requirements..161
Figure B-1 Example of DOF for a simple plate...164
Figure B-2 Six degrees of freedom at one node..164

xiv

Table of Tables:
Table 1-1 Flexible assembly tolerance analysis steps......................................................... 4
Table 1-2 Input/output summary of process steps ............................................................. 6
Table 2-1 Mean & standard deviations equations for 1-D assembly stackup................... 13
Table 3-1 Numerical derivatives [Gao 1998]. .................................................................. 24
Table 3-2 Rotational and translational variations associated with geometric feature
tolerance for 3D Planar and Cylindrical Slider joints [Chase 1996] ........................ 29
Table 3-3 Tolerance sensitivity matrices for determined and over-determined systems.. 32
Table 4-1Tooling/fixture dimensions for example assembly ........................................... 55
Table 4-2 Part dimensions for example assembly ............................................................ 56
Table 4-3 Summary of steps to determine misalignment ................................................. 69
Table 5-1 Summary of FEA modeling steps..................................................................... 80
Table 5-2 Material properties of parts A & B................................................................... 81
Table 5-3 Summary of steps to FEM compliant parts ...................................................... 89
Table 6-1 Ranges of surface variation and their effects ................................................. 101
Table 6-2 Surface continuity model use range .............................................................. 102
Table 6-3 Surface continuity models for example assembly .......................................... 107
Table 6-4 Summary of steps to calculate gap covariance............................................... 111
Table 6-5 Summary of steps for statistical FEA solution ............................................... 124
Table 7-1 Geometric tolerance variable for groove and cylinder example .................... 129
Table 7-2 Outline of FASTA inclusion of form variation at flexible joints ................... 132
Table 9-1 Strengths and limitations of FASTA .............................................................. 153
Table A-1 Vector loop modeling rules [Trego 1993]..161
xv

xvi

CHAPTER 1

Introduction

1.1 Introduction
In the fiercely competitive global economy of today, product success is awarded
to the company that can guarantee quality at a low cost. With such demands, the use of
tolerance analysis has become a vital link between product design and manufacture. By
nature, manufacturing is not an exact science. Tool wear, fixture imperfections, chatter,
and countless other causes produce deviations from a parts ideal design. When
designers call out nominal dimensions, they must also make an allowance for variations
in the actual parts made. Such deviations are permitted, within limits, through the use of
tolerances. The challenge is to permit as much variation as possible, to minimize
production costs, while still meeting critical functionality when components are
assembled. Tolerance analysis takes the guesswork out of tolerance assignment by
determining how sensitive the critical assembly dimensions are to part variation (see
Chase & Parkinson 1991).
Although much work has been done in this field, conventional tolerance analysis
methods are limited by the assumption of part rigidity. In the case of flexible parts,
standard tolerance analysis methods are unable to predict the effects of part compliancy
in an assembly. Such assemblies are often force-closed and then fastened, while the
flexible parts are held in a deformed state. This results in residual stresses and spring1

back deformations which can shorten product life and cause poor fit and poor
performance. An example of this type of assembly is illustrated in Figure 1-1, where
non-ideal parts A and B are forced together and fastened in a deformed ideal position.
Part c of the figure shows how part spring-back affects the final state of the assembly.
Part A

Gap

Part B
a) Compliant parts A & B to be assembled.
FA
FB
b) Forces used to place parts for fastening.

c) Resulting deformation when forces released.


Figure 1-1: Illustration of flexible assembly deformation

A system for statistical tolerance analysis of flexible assemblies would allow for
statistical prediction of the closure forces, deformations and residual stresses based upon
part tolerances. The need for this analysis method can be seen in the aerospace and
automotive industries, which routinely use flexible parts in production. Despite the
demand, there have not been, until recently, efficient means of characterizing compliant
assemblies.
New methods of performing tolerance analysis on flexible part assemblies have
been proposed at Brigham Young University, using Finite Element Modeling (FEM) and
statistical tolerance analysis. While these new methods have been verified by simple
simulation, they have not yet been used on a real multiple part assembly problem. To
2

predict the effects of the assembly gaps at the point of compliance, rigid body tolerance
analysis must be combined with flexible tolerance analysis to characterize the final state
of the assembly. The aim of this thesis is to present a comprehensive, integrated
methodology for this combined analysis, and to illustrate the process using a real
assembly.

1.2 Overview of New Methods


The challenge of statistically modeling an assembly with flexible parts is to
accurately take into account the effects of part compliancy. New methods have been
developed at Brigham Young University using a combination of Statistical Tolerance
Analysis (STA) and Finite Element Analysis (FEA) to achieve compliant assembly
tolerance analysis. Covariance, the interdependence of each element variation upon the
other elements, has been a key factor in effectively modeling compliant part behavior.
Internal material interactions, surface continuity constraints, and surface variation
wavelength all play a part in modeling flexible part covariance.
Statistical tolerance analysis of assemblies with compliant parts can be broken
down into four tasks as presented in Table 1-1. The process containing these four steps
will be referred to in this thesis as Flexible Assembly Statistical Tolerance Analysis, or
FASTA for short.

Table 1-1 Flexible assembly tolerance analysis steps

Summary of Analysis Steps


1.

Determine Misalignment (conventional STA)

2.

Finite Element Modeling of Compliant Parts


(FEM)

3.

Covariance Calculation (COV)

4.

Statistical FEA Solution for mean and variance


of Assembly Forces, Deformations and Stresses

Step 1: Determine Misalignment (STA). This entails the use of conventional


rigid body STA to solve for gaps between parts that result from variation stack-up in
assembly. Compliant parts are assumed to be rigid at this point, allowing for
conventional tolerance analysis. Inputs to this step are part geometry and part variation,
while outputs include the mean { i } and variance of every fastening point i along the
compliant parts at each gap. The variance is separated into translational, rotational, and
2
surface components ( { Ti2 }, { Ri2 }, { SPi
}) for use in step 3. Though suggested before, the

general adaptation of conventional STA for use on compliant assemblies has not
previously been done and thus is a major contribution of this thesis.
Step 2: Finite Element Modeling of Compliant Parts (FEM). This step accounts
for part flexibility in the overall analysis. The internal elastic material interactions
(Material covariance) of a compliant part are accounted for by using FEM to
approximate a parts material stiffness [K eq ] . Using sub-structuring techniques the
resulting stiffness matrix can be reduced to include only the degrees of freedom at
4

boundaries while still accounting for internal forces. This reduction in stiffness matrix
order allows for increased analysis efficiency. Inputs to the analysis include part
geometry and material properties. A significant contribution of this thesis is a general
method for performing this step using ANSYS, a widely available FEA software
package.
Step 3: Covariance Calculation (COV). The effects of surface continuity on the
mating surfaces of flexible parts are modeled statistically by the Geometric covariance.
To calculate the geometric covariance, sinusoids and polynomial series are used to model
continuous surfaces and solve for the covariant sensitivity, which constrains the mating
surfaces to be continuous. Spectral analysis methods can also be used to include surface
waviness effects in the covariance. A hybrid method allows for joining multiple surface
modeling techniques to calculate geometric covariance when no single model adequately
represents a particular surface variation.
The variance components of the assembly gap are used in conjunction with a
surface continuity model to solve for the geometric covariance of mating gap surfaces
and create a gap covariance matrix solution [ ] .
Step 4: Statistical FEA. Finally, the statistical FEA solution is calculated using
the outputs of the other steps. Using the equivalent stiffness [K eq ] , the mean of the
assembly gap { i } , and the gap covariance [ ] , the mean and the variance of the force
F( , ) is calculated. Once the statistical variation of the assembly surfaces has been
determined, the mean and variance of the displacement and stresses seen throughout the
compliant parts can also be determined by applying the resultant mean and variance of
the assembly gap as displacement loads in a FEA solution.
5

The product of these four steps is a means of predicting the closure force, deformation,
and stress throughout a flexible assembly over a statistical range. All of this is done with
minimal computation, making it much more efficient than simulation methods, such as
Monte Carlo Simulation (MCS). A summary of the inputs and outputs of each of the
process steps is presented in Table 1-2.
Table 1-2 Input/output summary of process steps

Process Steps
1. STA
2. FEM
3. COV
4. Stat. FEA

Inputs
Part Geometry
Part Variation
Part Geometry
Material Properties
2
{ Ti2 }, { Ri2 }, { SPi
}
Surf. Cont. Model
[K eq ]
{ i }
[ ]

Outputs
2
{ i } ; { Ti2 }, { Ri2 }, { SPi
}
[K eq ]
[ ]
F( , )
( , )
Stress( , )

While previous work has been done in each of the steps presented, a general
method for linking them together into a single process was never achieved. The product
of this thesis is a completely integrated FASTA process for general use on flexible
assemblies. A process flow diagram for this integrated process is shown in Figure 1-2.
Each step is shown as a part or module of the overall method, and is denoted by a boxed
section of the process. As each step is presented in more detail in the course of this
thesis, Figure 1-2 should be used as a reference frame or map to understand how it fits
into the complete FASTA process.

Figure 1-2 Process flow diagram of FASTA process

1.3 Thesis Objectives


The objectives of this thesis are: to include rigid part variation in the statistical
tolerance analysis of flexible assemblies, to present a general step-by-step methodology
for the FASTA process, and to illustrate how to apply the process using a real assembly
from industry. Significant contributions are made in the modeling of assembly
misalignment, and part compliance using FEA (steps 1 and 2 in Table 1-1).

1.4 Procedure
In order to achieve the objectives of this thesis the chapters are presented as
follows:
Chapter 2:

Background research of related topics is presented.

Chapter 3:

An introduction to rigid body statistical tolerance analysis is given

to set the foundation for FASTA.


Chapters 4-6: These chapters focus in detail on the 4 steps listed in Table 1-1.
The first step is treated in chapter 4, the second in chapter 5, and the third and
fourth in chapter 6. An example assembly is used to illustrate the modeling
procedures and analysis results at each step. Other considerations and information
are presented as relevant to each chapter.
Chapter 7:

Form variation in flexible assemblies is treated in this chapter.

Considerations for the modeling of mating flexible joints and variation


propagation through multiple joints in assembly are discussed.
8

Chapter 8:

The process presented in chapters 4-6 is applied to a real assembly

obtained from industry. Inputs needed for analysis are highlighted, and
measurement techniques for analysis and verification of the process are discussed.
Chapter 9:

Conclusions regarding the overall methodology are presented as

are recommendations for future work.

Appendices containing relevant data, more detailed examples, and modeling tutorials are
located at the end of the chapters.

1.5 Scope Delimitations


The analysis presented in this thesis will be limited by the assumptions and
conditions presented in the following sub sections.

1.5.1 Linear analysis


Linear assumptions limit the use of the methods contained in this thesis to parts
containing small variations. However, in tolerance analysis only small variations are
generally seen, making this limitation acceptable. Small elastic deformations are also
required so that compliant parts may be modeled at nominal dimensions, and yet still
accurately describe a statistically varying set of similar parts.

1.5.2 Statistical assumptions


Variation stack-up in assembly is assumed to be normally distributed. As the
combined effect of independent but non-normal variables tends to be normal, this usually
is a valid assumption. Tolerance values are assumed to represent three standard
9

deviations from the nominal part dimensions. This agrees with common industry
practice.

1.5.3 FEA modeling


Access to finite element models of compliant parts or FEA software to create such
models is assumed. It is also assumed that part stiffness matrices can be exported from
FEA models. ANSYS version 5.7 will be used for all examples and applications in this
thesis.

1.6 Summary
The product of this thesis is a step-by-step method to perform Flexible Assembly
Statistical Tolerance Analysis (FASTA), including variation from both rigid and
flexible parts. This combined analysis method will provide a new flexible assembly
analysis tool for industry, allowing designers to:

Predict the effects of variation on assembly processes.

Predict the range of forces required for assembly.

Predict the range of resultant stresses and deformations.

Estimate the effects of variation on assembly life, appearance, performance,


and additional labor and rework.

Evaluate design improvements.

Identify assembly problems, chief variation contributions, and focus process


control efforts.

It is hoped that such a tool may prove invaluable in the efforts of industry to provide
quality products at low prices, through the aid of statistical tolerance analysis.

10

CHAPTER 2

Background Research

2.1 Background
Pertinent to the scope of this research is a background in both rigid and flexible
statistical tolerance analysis. Brief introductions to these will be made in this chapter,
while a more detailed presentation of work foundational to this thesis is presented in
relevant chapters.

2.1.1 Tolerance analysis of rigid assemblies


Much work has been done in the realm of tolerance analysis for rigid parts. As
discussed earlier, the premise of this analysis is that assembly variations result from a
stackup or accumulation of component variations. Thus, the assembly variations are
dependant upon size and form variations in the parts.
Three analytical models commonly used to predict the accumulation of
component tolerances in an assembly are: Worst Case, Statistical, and Monte Carlo
Simulation. Their definitions and characteristics are as follows:
1. Worst Case (WC) assumes that all part dimensions are simultaneously at
their worst tolerance limit. Although sizing tolerances by worst case assures
that all assemblies will be within the required limits, it does so by demanding
11

tight tolerances at increased cost. Since the chance of all part dimensions
being at their extreme limits is very small, this cost may be excessive.
2. Statistical or Root Sum Squares (RSS) methods use statistical distributions to
represent the probable magnitude of variation in manufactured parts.
Tolerances are generally assumed to correlate to six standard deviations (6
or 3). This method allows for relaxed tolerances compared to WC, at the
expense of some rejects.
3. Monte Carlo Simulation (MCS) uses random perturbations of part
dimensions within tolerances to simulate a population of complete assemblies.
In this way, assembly tolerances can be checked by a simulated production
run. This method is limited by the computational cost required to simulate
large populations (typically 1000 or more), particularly when each simulation
requires a FEA solution.
For each of these, the mean and standard deviation equations are shown in Table 2-1.
Statistical methods are preferred for flexible assembly applications because they
realistically model part variation and are computationally efficient.

12

Table 2-1 Mean & standard deviations equations for 1-D assembly stackup

Model

Mean

Standard Deviation

WC

x ASM : Mean assembly

x ASM = xi

T ASM = Ti

i =1

dimension
x ASM : Measured assembly
dimension
xi : Mean dimension

i =1

RSS

Where:

x ASM = xi

ASM

i =1

T
= i
i =1 3
m

xi : Measured dimension
ASM : Standard deviation
Ti : Tolerance value

x ASM j = xi
i =1

MCS
x ASM =

x
j =1

ASM

(x
n

ASM =

j =1

ASM

x ASM

n 1

m:

Number of parts in
assembly

Number of
n : simulated
assemblies

In addition to analysis through simulation, as in Monte Carlo, a linearized


statistical model has been developed for 2-D and 3-D tolerance analysis called the Direct
Linearization Method (DLM) [Chase 1997]. Vector loops are used in this method to
represent the 2-D and 3-D dimensional chains in an assembly that cause tolerance
stackup. Resulting nonlinear equations are linearized by using the first-order Taylor
series. This simplification assumes small variations from nominal dimensions, a very
reasonable assumption, since tolerances are generally orders of magnitude smaller than
associated dimensions. Tolerance sensitivities are calculated using partial derivatives of
the vector loop along coordinate axes. The linearized loop equations are put into matrix
form to calculate predicted assembly variation. Dimensional, form, and kinematic
variations can all be included in the analysis, making it quite versatile.
In complex 3-D assemblies, obtaining tolerance sensitivities through partial
derivatives can be tedious and confusing. [Gao 1998a] presented a simpler way of
13

calculating these sensitivities called the Global Coordinate Method (GCM). Using vector
loops and the relative lengths and rotational adjustments between vectors, closed-form
sensitivities are derived by approximating the effects of small perturbations on the loop
closure equations. The sensitivity of the outputs can be inserted directly into the DLM
for use in the matrix algebra that completes the calculation of tolerance sensitivities.
Only a brief introduction to rigid body tolerance analysis has been presented.
More detailed information will be presented in Chapter 3 for methods foundational to this
thesis. For a more complete survey of tolerance analysis pertaining to rigid assemblies
see: Chase & Parkinson (1991), and Narahari (1999).

2.1.2 Tolerance analysis of flexible parts


The effects of part compliancy on assembly tolerances are a relatively new area of
study, with significant analytical work beginning in the mid 1990s. Since there are as of
yet no standard methods to this analysis, background on this topic will be presented as a
review of contributing literature.
Dr. Jack Hu and his students at the University of Michigan have made significant
contributions to the analysis of sheet metal assembly variation. [Liu 1995] presented a
one-dimensional offset beam element model used to evaluate the affects of part variation,
tooling variation, and assembly sequences by Monte Carlo Simulation. Assuming linear
mechanics, [Liu 1996] reviewed the effects of sheet metal assembly sequence and the
effects of multiple joints using simple examples. [Liu 1997] derived influence
coefficients for sheet metal assemblies using linear assumptions. This minimized the

14

number of calls to the assembly Finite Element Model (FEM) during MCS, markedly
improving analysis efficiency.
Also from the University of Michigan, [Shiu 1997] offered a procedure of
modeling the sheet metal in automotive body structures as flexible beams. The simplified
beam model was then used for dimensional control in the assembly. While a beam
structure model is less complex than a full FEA model, it appears that the full FEA model
is still required for beam structure derivation. This modeling redundancy makes the
method less valuable.
[Chang 1997] modeled the assembly process of non-ideal compliant parts, in a
single assembly, using fastening and measurement cycles. Contact chains were used to
represent the interactions between parts and tooling in these cycles. This model was used
to simulate the propagation of variation in the assembly process and to predict variations
in the resulting assembly. Statistical results were not discussed.
From France, [Sellem 1998] looked into the feasibility of integrating FEM/CAE
analysis on deformable assemblies. Thin shell theory was used to assume linear
elasticity, and three influence coefficient matrices were implemented for variation in part
positioning, tooling conformity, and shape. [Sellem 1999] compared analysis predictions
using shape variation with measured assemblies for validation. The simulated results had
the same global behavior for the mean and range of the measured inspection point
deviations. Statistical predictions were not incorporated into this model.
[Hockmuth 1998] suggested that elastic deformation of mating parts be accounted
for by using a FE-mesh to calculate stiffness matrices. It was also suggested that a parts

15

material covariance at contact be taken into account using a partitioned stiffness matrix,
though no application of this idea was presented.
[Hu 2001] presented a numerical simulation method to analyze the assembly
process of compliant parts. ANSYS FEM code was used to combine elastic and contact
analysis to model part compliance and assembly. Part variation, assembly tooling
variation, and welding distortion were modeled in this simulation non-statistically.
At Brigham Young University (BYU) a group of students advised by Dr. Ken
Chase have developed a methodology whereby a finite element model is used in
conjunction with statistical analysis methods to statistically describe flexible part
assembly processes. The FASTA method (presented in Chapter 1) permits the prediction
of the range of assembly forces and the resulting stresses and deformations in the
assembly due to dimensional and form variation. Only two solutions are required to
characterize the variation in an entire production run. This methodology has evolved
through a compilation of work to include the affects of material interactions, surface
continuity, and surface waviness. The inclusion of these affects allow for a more
accurate analytical model of compliant part assembly processes.
Random statistical methods alone are unable to describe elastic material
coupling or surface continuity effects on the deformation of mating surfaces during
assembly. [Merkley 1996,1998] solved this problem using statistical covariance, the
interrelation of variation at one point to another. Internal material interactions (Material
Covariance) were derived from the material stiffness matrix (Keq) obtained from FEM.
Surface continuity constraints (Geometric Covariance) were included by generating
random input surfaces rather than random points to simulate mating surface errors. Such
16

random surfaces can then be used to derive a sensitivity matrix used in the analysis to
include geometric covariance. A method involving Bezier curves was suggested as a
means for creating the random input surfaces. Merkley also applied higher order
elements (Super-Elements) to condense each parts stiffness matrix and improve analysis
efficiency.
[Stout 2000] focused on simplifying the calculation of geometric covariance,
suggesting a method that uses nth order polynomials to approximate random continuous
input surfaces. It was also discovered, through simulation, that surface waviness on
mating surfaces has a significant effect on the resultant internal stresses of assembled
flexible parts. Both the Bezier curve and polynomial methods do not account for this
additional stress.
[Bihlmaier 1999] used spectral analysis of surface profiles and Fast Fourier
Transform (FFT) techniques to obtain the geometric covariance directly from the
frequency spectra of mating surfaces. This was done in order to include the affects of the
wavelength of surface waviness in flexible tolerance analysis. [Soman 1999] developed a
method for characterizing surface errors statistically so that real part data could be used
as input to a statistical FEA of assemblies. He statistically combined surface scans of
sheet metal parts and performed Fast Fourier Transforms (FFT) on the measured data to
obtain input frequency spectra. It was determined that spectral analysis using Discrete
Fourier Transforms (DFT) is limited to surface wavelengths less than or equal to the
mating surface length.
[Tonks 2002] continued the work of Stout and Bihlmaier by developing a hybrid
method to combine surface continuity models in the calculation of geometric covariance.
17

He also derived a means of modeling continuous surfaces using sums of sinusoids. This
allows for the analysis of non-integer surface wavelengths, as measured or predicted from
part surfaces (overcoming a limitation of DFT spectral analysis). Using complete
orthogonal polynomial series he was also able to accurately model geometric covariance
for surface wavelengths greater than the part length, another limitation of spectral
analysis.
A primary contribution of Dr. Chase and his students to the proper modeling of
variation in flexible assemblies is the inclusion of geometric covariance. In summary, the
six methods of accounting for covariance of continuous surfaces are:
1. Monte Carlo generates sets of random surfaces.
2. Polynomial creates a covariance matrix using a sensitivity matrix
obtained from a polynomial curve fit.
3. Bezier creates a covariance matrix using a sensitivity matrix obtained
from a Bezier curve fit
4. Spectral Analysis creates a covariance matrix using the average
frequency spectrum obtained from a FFT of the mating surfaces.
5. Sinusoidal creates a covariance matrix using a sensitivity matrix
obtained from a sinusoidal curve fit.
6. Hybrid Analysis creates a covariance matrix using a combination of
other methods based on weighting factors.

2.2 Summary
Only an overview of the literature reviewed has been presented in this section.
This thesis will primarily focus on and add to research performed at Brigham Young
University. More detail will be given for research relevant to specific topics in
subsequent chapters.
18

CHAPTER 3

An Introduction to Rigid Body


Tolerance Analysis

3.1 Introduction
Before presenting a methodology for the statistical tolerance analysis of flexible
assemblies, it is useful to first understand the fundamentals of rigid body tolerance
analysis. In this way it is possible to start with a solid foundation in modeling assembly
variation propagation between rigid parts, allowing for later adaptation to include flexible
parts. With this in mind, the purpose of this chapter is to provide a general introduction
to rigid body tolerance analysis principles and procedures. Sources of assembly variation
are first discussed, and then vector loop modeling techniques (specifically the Direct
Linearization Method) are presented to account for such variation.

3.2 Sources of Assembly Variation


In order to understand assembly variation, one must look at the individual
components in an assembly. Deviation of individual parts from ideal or perfect
dimensions is a result of imperfections in manufacturing processes. As a result,
dimensional and form variations of manufactured parts accumulate statistically and
contribute to variation in critical assembly dimensions. This stack-up of dimensional and
form variation is facilitated by kinematic adjustments between parts during assembly
19

[Chase 1997]. Tolerances are used as a means of limiting deviations from desired
geometry. Dimensional tolerances control part sizes by prescribing acceptable variations
in lengths and angles. Form tolerances are used to control feature variations such as
flatness, roundness, angularity, etc. as defined by ANSI Y14.5-1994 standards [ASME
1994].
The dependency of kinematic variation on dimensional and form variation is
illustrated in Figure 3-1. In part a, the assembly dimension U is dependant upon the
component dimensions A, R, and . Nominal part dimensions place the cylinder in the
groove with a center distance of U1, while variations from nominal change the placement
to U2. The difference between U1 and U2 is the kinematic adjustment of the assembly to
variation in the components. In part b of the figure, it can be seen that U is also
dependant upon form variation of the grooves surface.

A + A

R + R

U1

U
U2

a) Kinematic adjustments due to dimensional


variation

b) Kinematic adjustments due to


form variation

Figure 3-1 Kinematic Assembly Variation [Chase 1996]

20

While the sources of variation are nearly limitless, their results can be represented
as either dimensional, form, or kinematic variations. The tools required to model each of
these variations will be given in the following section.

3.3 Vector Loop Modeling


As explained in chapter 2, vector loops can be used to represent the dimensional
chains in an assembly that contribute to tolerance stack-up. A simple vector loop model
of the assembly presented in the last section is
U1

shown in Figure 3-2 as an example.


R

The Direct Linearization Method (DLM)

U2

makes use of vector loop modeling techniques to


1

calculate the sensitivity of an assembly to part


variations. Through linearizing assumptions and

Figure 3-2 Simple Vector Loop Model

matrix algebra, the DLM can be used to analyze

even complex assemblies. In this section the DLM is first introduced for 3D assemblies
with dimensional and kinematic variation. The inclusion of geometric form variation in
this method is then explained and recommended vector loop modeling practices are
presented.

3.3.1 DLM in 3D with dimensional and kinematic variation


The Direct Linearization Method for 3D tolerance analysis was presented by
[Chase 1996, 1997], and [Gao 1998b]. In this method, assembly constraint equations are
formed, using closed and open vector loops, to calculate the affects of variation in
21

component dimensions on the final assembly. Closed vector loops begin and end at the
same point, and are used to solve for kinematic variations. Open loops begin and end on
opposite sides of a clearance or other feature, allowing for the calculation of variation in
critical assembly features.
When dealing with variation in three dimensions, the most general case, resulting
vector loop constraint equations become increasingly complex. The transformation in a
vector loop from joint i-1 to i (between mating parts) includes a combination of four
matrices, one for translation and three for rotation. These 3D transformation matrices are
shown in equations 3-1 to 3-4, where x , y and z represent rotations about relative axes
and Tx, Ty, and Tz are translation vector components between consecutive nodes in a loop.
0
1
0 cos
x
[R x ] =
0 sin x

0
0

[R ]
y

cos y
0
=
sin y

cos z
sin
[R z ] = z
0

0
sin x
cos x
0

0
0
0

(3-1)

0 sin y
1
0
0 cos y
0
0

0
0
0

(3-2)

0
0
0

(3-3)

sin z
cos z
0
0

0
0
1
0

22

1
0
[T] =
0

0
1
0
0

0 Tx
0 T y
1 Tz

0 1

(3-4)

The constraint equations for a 3D vector loop can be written as a string of transformation
matrices multiplied together, as shown in equation 3-5. Here [Ri] is the product of
rotation matrices at joint i, [Ti] is the transformation matrix at i, and [Rf] is the final
rotation in the loop to bring it to closure.

[R 1 ][T1 ][R 2 ][T2 ]L[R i ][Ti ]L[R n ][Tn ][R f ] = [H ]

(3-5)

For closed loops the resulting matrix [H] is equal to the identity matrix, while for open
loops it represents a clearance or gap in the assembly.
As can be expected, the constraint equations for a 3D vector loop are nonlinear,
making them difficult to solve. In the DLM, however, small variations in component
dimensions allow for a good approximation of the solution using a Taylor expansion of
equation 3-5. In this way, only the derivatives are needed for the analysis. To obtain
these derivatives, small perturbations L and are introduced one at a time into
equation 3-5, at the point where variation occurs. The perturbations imposed on the loop
prevent it from closing and result in a closure error vector {X Y Z 1} in place of
T

[H]. By varying each length and angle separately in the matrix constraint equation, the
derivatives may be approximated numerically as shown in

Table 3-1 below. In these

equations Hx, Hy, and Hz are the translation constraints in the global x, y, and z axis; H x ,

23

H y and H z are the rotational constraints about the respective global axis; and finally
, and are the global direction cosines of the local axis of rotation i [Chase 1996].
Table 3-1 Numerical derivatives [Gao 1998].

Translational Variable
H x X

Li
Li
H y
Li

Rotational Variable
H x X

i
i
H y

Y
Li

Y
i

H z
Z

Li
Li

H z
Z

i
i

H x
=0
Li

H x
= cos
i

H y

H y

Li

=0

H z
=0
Li

= cos

H z
= cos
i

3.3.1.1 Closed vector loops


From the first-order Taylor expansion, the constraint equation for a closed loop
can be written in matrix form as:

{H } = [A]{X } + [B]{U } = {0}

(3-6)

where {H } : vector of clearance variations,

{X } :

vector of manufacturing variations or tolerances,

{U } :

vector of variations of the assembly variables,

[A] :
[B] :

matrix of partial derivatives with respect to manufacturing variations,


matrix of partial derivatives with respect to assembly variations.
24

The matrices in equation 3-6 are populated using the partial derivatives shown in
Table 3-1, with each column of [A ] containing the following format:

{Ai } = H x
xi

H y
xi

H z
xi

H x
xi

H y
xi

H z

xi

(3-7)

In equation 3-8, xi refers to the ith-manufactured dimension, meaning that there will be a
column for every manufactured dimension in the vector loop. Matrix [B ] has the same
column notation, but uses the assembly dimension variable ui instead of xi. When setting
up these matrices it is important to know which dimensions in the vector loop are
independent and which are dependent. Independent manufactured variations are
described in [A ] , and dependent assembly variations are described in [B ] .
With {H } = {0} in a closed vector loop, equation 3-7 can be solved in terms of
assembly variation {U } . For an assembly where matrix B is square, and nonsingular, a
determined system, this results in the following equation:

{U } = [B]1 [A]{X }

(3-8)

If matrix B is not square, meaning that the assembly vector describes an over-determined
system, then the least square fit solution is:

{U } = ([B]T [B]) [B]T [A]{X }


1

(3-9)

From equations 3-8 and 3-9 we can see that the assembly variations U can be obtained
from the component variations X by simple matrix algebra, once the sensitivity
matrices are known. The correlation or sensitivities between the two types of variation
will be discussed later.
25

3.3.1.2 Open vector loops


In an open vector loop the constraint equation is not equal to zero, instead
representing a clearance in the assembly. This often means that a closed loop solution
must first be solved to determine {U } , which can then be used in the open loop
constraint equation to solve for {V }:

{V } = [C]{X } + [D]{U }

(3-10)

where {V }: vector of variations of open loop assembly variables,

[C] :

matrix of partial derivatives with respect to open loop manufacturing


variables,

[D] :

matrix of partial derivatives with respect to open loop assembly variables.

If [B ] is a square matrix, then equation 3-9 can be inserted into equation 3-10 and
written as:

{V } = ([C] [D][B]1 [A]){X }

(3-11)

otherwise,

{V } = [C] [D][( B]T [B]) [B]T [A] {X }


1

(3-12)

Once again, in equations 3-11 and 3-12, it can be seen that assembly variation (in this
case representing a gap) can be determined using manufacturing variation and matrix
algebra, once the sensitivity matrices are known.

26

3.3.2 Including form variation in 3D DLM


Having presented the DLM equations for determining assembly variation using
dimensional and kinematic variations, it is not difficult to expand the method to include
form variation (also known as geometric feature variation). Form variation propagates in
assembly only through the surface contacts or joints between parts. This is modeled in a
vector loop at the point of contact using a zero-length vector, allowing for an independent
variation source whose mode of variation reflects the joint type between mating surfaces.
While there are many different types of mating contacts possible between parts,
discussion will be limited to variations resulting from planar and cylindrical-slider joints.
For a complete presentation of geometric feature variation in tolerance analysis, including
all joint types, see [Chase 1996].

3.3.2.1 Form variation at Planar and Cylindrical Slider joints


The effect of form variation on assembly is dependant upon the type of joint
between mating parts, and the joint axis under consideration. Based on these
considerations, feature variation can either result in rotational or translational variations
transmitted through mating parts. This is illustrated in Figure 3-3 using a 3D cylindrical
slider joint. In part A of this figure, it is apparent that surface variation within the flatness
tolerance zones of either part can result in rotation variation normal to the cylinder axis at
assembly. Looking down the cylinder axis (part B), it can be seen that variation within
the form tolerance zones results in translational variation normal to the contact surface.
This type of 3D variation is typical of parts that have a line contact on a surface.

27

y
Rotational Variation

Planar
Tolerance
Zone

A) View normal to the cyl. axis

3D cylindrical slider joint

Nominal
Circle
Translational
Variation

Cylindrical
Tolerance
Zone
Tolerance
Zone

The effect of feature


variations in 3D depends
upon the joint type and which
joint axis you are looking
down.

B) View looking down the cyl. axis


Figure 3-3 3D Effects of form variation [Chase 1996]

Another common assembly interface is a planar joint. Just as the name suggests,
planar joints have a planar contact between mating parts. A simple example of this type
of joint would be two blocks stacked together, where surface variation of either part
would result in rotational variation along either planar axis (as in part A of Figure 3-3).
In order to calculate the range of angular variation caused by variation along a
planar axis, the tolerance zone and the characteristic length (the length of contact between
the two surfaces) must be known. This variation can then be calculated using equation 313.
Tolerance Zone

= tan 1
Characteristic Length

( 3-13)

Translational variation normal to the contact surface, as seen between a line or point
contact between parts, is equal to 0.5 , where is the width of the tolerance band.
28

The degrees of freedom of kinematic motion (K) and geometric feature variation
(F) for cylindrical slider and planar joints are shown in Figure 3-4. It can be seen in this
figure that when K is removed by constraints, F has the possibility of propagating in an
assembly.

K Kinematic Motion
F Geometric Feature Variation
F

K
y

Cylindrical Slider Joint

Planar Joint

Figure 3-4 Degrees of freedom for kinematic motions and geometric feature variations
[Chase 1996]

The variations resulting from specific geometric feature tolerances for these two joints
are shown in Table 3-2. In this table, R denotes rotational variation and T represents
translational variation with respect to a referenced joint axis.
Table 3-2 Rotational and translational variations associated with geometric feature tolerance for 3D
Planar and Cylindrical Slider joints [Chase 1996]

Geometric Characteristic Symbol (Tolerance)


Joint
Planar

Rx Rz

Rx
Rz

--

--

Rx Rz

Rx Rz

Rx Rz

Rx Rz

Rx Rz
Ty

--

--

Ty
Rx

Ty Rx

Ty Rx

Ty Rx

Ty Rx

Ty Rx

Ty Rx

--

Ty

--

--

Ty Rx

Ty Rx

Ty Rx

Ty Rx

Ty Rx

--

--

Cyl. Slider
Cylinder
Plane

-Ty Rx

Ty
Rx
Ty
Rx

29

3.3.2.2 Dimensional, kinematic, & form variation in DLM


Once the form variation components have been determined, as presented in the
previous section, they can now be included in the DLM. With the geometric feature
variation of the assembled components, the linearized constraint equations become for
closed loops:

{H } = [A]{X } + [B]{U } + [F]{ } = {0}

(3-14)

And for open loops:

{V } = [C]{X } + [D]{U } + [G ]{ }

(3-15)

Where { }: vector of variations of form variables,

[F] , [G ] : matrices of partial derivatives with respect to closed and open loop
form variables.
Once again, {U } can be solved for in the closed loop equation (3-14) resulting in:

{U } = [B]1 [A]{X } [B]1 [F ]{ }

(3-16)

Equation 3-16 can then be substituted into equation 3-15 to obtain:

{V } = ([C] [D][B]1 [A]){X } + ([G ] [D][B]1 [F]){ }

(3-17)

These two final equations describe the variations of the assembly variables for closed and
open loops with respect to dimensional and form variation of individual parts. With these
equations, it is now possible to estimate the kinematic variations.

30

3.3.2.3 Sensitivities and kinematic variation estimates


In equations 3-16 and 3-17, the relation of dependent variables to independent
variables is a product of matrix algebra that forms tolerance sensitivity matrices where:
[S d ] represents the tolerance sensitivity matrix for dimensional variables, and[S ] for
form variables. For open and closed loops the tolerance sensitivity matrices for
determined and over-determined systems are summarized in Table 3-3.
Using these sensitivity matrices, the kinematic assembly variations can be
estimated using a tolerance accumulation model. Worst Case and Statistical (R.S.S)
accumulation models are shown below.
Worst Case Model:
n

j =1

j =1

U i = S ijd tol ijd + S ij tolij Tasmi

(3-18)

Statistical Model:

U i =

(S
n

j =1

d
ij

tol ijd

) + (S
2

j =1

ij

tol ij

Tasmi

where: tolijd : vector of tolerances for dimensional variables,


tolij : vector of tolerances for form variables,
U i : vector of kinematic variation estimates (assembly variations),
n: number of dimensional variables,
m:

number of geometric variables,

Tasmi : design limit for assembly variation ofU i .

31

(3-19)

Table 3-3 Tolerance sensitivity matrices for determined and over-determined systems

System
Determined:
[S d ] =
[S ] =
Over-Determined:
[S d ] =
[S ] =

Closed Loop

Open Loop

[B ] [A ]

[C] [D][B]1 [A]


[G ] [D][B]1 [F]

[B] [F ]
1

(
)
([B ] [B])

[B] [B]
T

[B]T [A]
1
[B]T [F]

[C] [D][( B]T [B]) [B]T [A]


1
[G ] [D][( B]T [B]) [B]T [F]
1

Presented thus far are the tools needed to determine rigid assembly variation
based on dimensional and form variations, whether measured or predicted, using
tolerance values. Equations 3-18 and 3-19 provide the final step in modeling the stack-up
or sensitivity of assembly variations (for closed or open loops) to part variation. The
statistical accumulation model (equation 3-19) is a more realistic prediction of how part
variations actually accumulate in assembly and will be the only model used hereafter.
Worst case accumulation (equation 3-18) is an overestimate of the assembly variation,
and has application when stringent requirements are demanded of critical assembly
features. Such requirements are often called for when designing for military application.

3.3.3 Vector loop modeling practices


When modeling any physical system, it is important to remember that the analysis
results are only as good as the model used. In the case of vector loop modeling for
tolerance analysis, knowing how to set up the model can be the most challenging part. In
order to create valid vector loop models, rules and requirements have been set up to
check models for error. A list of general modeling rules for 2-D assemblies is presented
32

in Appendix A. These can be applied to 3-D vector loops by viewing each dimension as
a 2-D plane. If at all possible, vector loop models should be checked with measured data
for validation.
It should be noted that dimensioning schemes play a significant role in tolerance
analysis. Vector loops follow dimensional variables to model variation stack-up, and are
thus also dependent on dimensioning schemes. Datum positions and references do have
an impact on tolerance stack-up in assembly, and should thus be chosen carefully.
Tolerance analysis provides a means of comparing the assembly results of various
dimensioning schemes, allowing for an educated choice between options. It should also
be stated that some dimensioning schemes do not easily lend themselves to vector loop
modeling. As practice in this art is obtained, a designer will be able to plan ahead for
tolerance considerations and avoid later problems in production.

3.4 Summary
In this chapter dimensional and kinematic variations have been discussed and
modeled for rigid assemblies using the Direct Linearization Method. Form variation
propagation at planar and cylindrical slider joints was presented and also included in the
DLM. Sensitivity matrices were used in conjunction with accumulation models (worst
case and statistical) to predict variation of critical assembly features. Vector loop
modeling practices were also discussed. With this introduction to rigid body statistical
tolerance analysis, a foundation has been laid for determining misalignment in flexible
assemblies, the first step of FASTA and the topic of chapter 4.

33

34

CHAPTER 4

Determine Misalignment:
DLM for Flexible Assemblies

4.1 Introduction
When modeling the assembly of flexible parts, it is necessary to determine the
mean and the variance of the gap between mating surfaces at each assembly node. This
means that a vector loop model must be extended to include each point where fastening is
to occur (by weld, rivet, etc.). Once the gap variation is characterized statistically, the
range of variation can be used to determine the range of assembly force, residual stresses
and deformations for the whole production run. This chapter discusses how to apply the
DLM to flexible assemblies, introducing vector loop techniques and matrix notation. A
system for resolving gap variation into separate components is also described in order to
overcome modeling limitations encountered when calculating covariance (see Chapter 6).
The final process is illustrated with an example.
As discussed briefly in Chapter 1, conventional rigid body STA provides the
means to estimate variation in the gaps between parts that result from variation stack-up
in an assembly. By treating compliant parts as if they were rigid during this step
(modeled without an applied force), the mean and variance of the assembly gap can be
calculated strictly due to dimensional variation. Figure 4-1 shows a graphical overview

35

of the process used in step 1 of FASTA to statistically determine misalignment, including


both required inputs and solution outputs.

FASTA Step 1 Overview


Part Variation

Part Geometry

Inputs

1. STA

Gap Variation

{ } { }
2
Ti

2
Ri

{ }
2
SPi

Mean

{i }

Outputs

Figure 4-1 Step 1 process summary

As shown in this figure, the gap variation output is broken into three components
(translation, rotation, and surface variation) for use in subsequent steps of FASTA. The
mean gap vector is also determined in this step.
In this presentation a flexible or compliant part is defined as a thin-walled
component that can be bent or flexed elastically during the assembly process. This can
include sheet metal parts as well as composites or any other material with known
mechanical properties. Since the goal of the overall analysis is to determine the forces
required to join flexible parts in an assembly, it is useful to think of this step in the
overall process in terms of a linear force-deflection equation:

{F} = [K ]{}

(4-1)

36

Where in matrix form {F} represents the closure force at each fastening point, {} is the
closure distance or gap at each fastening point, and [K] is the stiffness of the compliant
part. In this context, the purpose of this chapter is to present a way to statistically predict
the variation in {}.

4.2 Modeling Flexible Assemblies Using DLM


The direct linearization method, presented in Chapter 3, allows for the inclusion
of multiple variation modes in the statistical analysis of variation stack-up in assembly.
With some minor adaptations, this versatile method can be used to successfully analyze
assemblies that contain both rigid and flexible parts.

4.2.1 Obtaining the mean gap solution


An important task of the statistical solution, determining the mean of the gap
value at each fastening point along a flexible joint { i }, is a required output of the first
step of the FASTA process. Generally, the mean separation between assembled flexible
parts is designed to be zero. However, imperfections in manufacturing processes can
introduce mean shifts that result in non-zero mean values. In some cases, it may even be
desirable to deliberately design for a mean separation in order to preload a joint using
forced flexure of compliant parts.
The easiest method for obtaining the nominal gap values at each fastening point is
to query CAD models of the assembly being analyzed. Most CAD packages include
tools for measuring assembly distances between parts. Since the parts are modeled
ideally with nominal dimensions, the output assembly distances represent nominal values.
37

In a parametric CAD system, the nominal dimensions can be replaced with actual mean
shifts to automate the solution of the mean assembly gap.
If CAD models are not available, kinematic assembly equations created in the
DLM method provide another option. For flexible assemblies that can be modeled using
only open vector loops, the mean gap is simply the solution to the kinematic equations
evaluated at nominal dimensions. If closed vector loops are required to model an
assembly, then iterative solutions are necessary to solve for unknown assembly
dimensions.

4.2.2 Modeling flexible gap variation


The first step in modeling a flexible assembly is to create a vector loop that
accounts for the chain of dimensions that contribute to each assembly variable. In this
case, the variation of a gap at each fastening node is the desired output of the analysis.
While a separate vector loop could be drawn for each fastening location, an assembly
containing a row of 500 rivets demands a more efficient model. Such a model can be
created using variable vectors along the fastening line, as shown in Figure 4-2.

38

M odeling M ultiple F astening P oints


U sing V ariable V ectors ( fi, gi)

Loop 2
f2
b

g2

G2
c

b
G1

Loop n
c

fn

Gn

Loop 1

gn
c

Figure 4-2 Modeling multiple fastening points using variable vectors (fi, gi)

In this figure, vector loop 1 is 2-dimensional and completely contained in the x-y plane.
This initial loop serves as a base-loop for the modeling of all of the fastening points; with
subsequent vector loops extending into the third dimension using variable vectors fi and
gi. By varying the length of these vectors, it is easy to quickly represent the open vector
loop for n fastening location in the assembly with just one vector loop model. It should
be noted that each vector chain is still treated as independent of the other loops since
vectors fi and gi are independent from point to point (for i = 1, 2, 3, . . .n). The
interdependence between each fastening point (due to surface continuity) is accounted for
in Chapter 6 using covariance models.
Form variation along fastening surfaces can also be modeled using zero-length
vectors (fi, gi) at each fastening node as presented in Figure 4-3. As shown in this
figure, the vectors are oriented to model the resultant tolerance zone (contained by dotted
lines). At this point only form variation of flexible parts along the gap under
39

consideration will be included in the analysis. Form variation at rigid joints can be
included, but surface variations at flexible joints other than the gap are neglected.
Chapter 7 discusses in detail the topic of form variation of flexible parts in assembly, but
the inclusion of such in FASTA analysis is beyond the scope of this work.
Form Variation at Gap
fi

fn
gn

f1

gi

g1

Figure 4-3 Form variation at multiple fastening points

Adjusting the DLM matrix notation and equations for use with a flexible
assembly containing multiple fastening nodes (using variable vectors) is not difficult.
From Chapter 3, the general DLM equation for an open loop is:

{V } = {Gmi } = [C]{X} + [D]{U } + [F]{ }

(4-2)

Where {Gmi }will be used to denote gap variation in the m direction at fastening node i.
In the most general solution, the subscript m includes the set of global coordinate
directions and rotations (x, y, z, x, y, z). If all six degrees of freedom are considered at
each fastening node, vector and matrix sizes can get very large (#rows = 6i). In many
cases, however, variation analysis can be limited to only one or two degrees of freedom
of interest. In this thesis, analysis will generally be limited to a single coordinate axis
which describes the direction of the force used to close a gap.

40

Including only the variation in the y-direction, the assembly shown in Figure 4-2
and Figure 4-3 can be treated as an open vector loop by solving for the unknown gap
vector in terms of the known part dimensions and variations. The general DLM vectors
and matrices can then be written as:
a
b

f 1

g 1
{X } = M
f
n
g n

d
e
G y1

a
G y 2
[C] = a

M
G yn
a

G y1

f 1
G y 2
[F] = f 1

M
G yn

f1

f 1

f1
1
{ } = M
2

fn
gn

G y1

G y1

G y1

b
G y 2

f 1
G y 2

g 1
G y 2

b
M
G yn

f 1
M
G yn

g 1
M
G yn

f 1

g 1

G y1
g1
G y 2
g1
M
G yn
g1

L
L

G y1
fn
G y 2

fn
O
M
G yn
L
fn

G y1

G y1

G y1

f n
G y 2

g n
G y 2

d
G y 2

f n
O
M
G yn
L
f n

g n
M
G yn

d
M
G yn

g n

L
L

G y1

e
G y 2
e

M
G yn
e

(4-3)

G y1

gn
G y 2

gn
M
G yn
gn

In matrices [C] and [F ] it is clear that each column describes the sensitivity of every
fastening point to a single dimensional or form variation. Thus, there are as many
columns in these matrices as there are respective variation components included in the
41

vector loop model. Missing from equations 4-3 are the closed loop components: [D]
and {U } . For the given assembly these have no values because there are no closed loops
required to model the assembly variation. When closed loops are required, [D] is similar
to [C] , but partial derivatives are taken with respect to assembly variables rather than
manufacturing variation. The assembly variation vector, {U } , also has the same form as

{X } .
Simplifications to the general matrices presented are possible in many cases. For
example: due to the independence of each variable length vector fi, G yi f j = 0 if i j .
The same is true for the partial derivatives of each gap with respect to vectors gi, fi and
gi. Also, since variable vectors fi and gi are opposite in direction they cancel each other
out nominally whenever they are parallel and equal in length. If the location of the
fastening points is located by a single process, such as drilling rivet holes on mating parts
in a fixture, then the variation of these vectors also cancels out. When these vectors are
orthogonal to the gap direction in question, then the partial derivative with respect to the
variable vector is zero, and they can be omitted from the analysis. Matrix [F ] is
simplified for a uniform tolerance band (meaning fi and gi are constant along the
surface), reducing the number of columns significantly.

4.2.3 Resolving gap variation into components


Gap variation in flexible assembly can be broken down into three types or modes:
translational, rotational, and surface variation. This separation of variation contributions
is illustrated for a single flexible part surface in Figure 4-4.
42

Translation
Single Part Variation
from Ideal
Rotation
actual

Surface

Ideal (mean)

Figure 4-4 Example of modes of variation contribution

Spectral analysis methods used to model covariance in step 3 of FASTA (see Chapter 6)
are unable to handle theses rigid body modes with wavelengths longer than the part
length. In order to overcome these limitations, the gap variation must be broken down
into the variation components shown in Figure 4-4 for individual treatment.
Additionally, resolving variation into separate components aids in the identification of
variation sources and the determining of dimensional percent contribution to assembly
variation. This information is useful for identifying chief variation contributions and
focusing process improvement efforts. In this section translational, rotational and surface
variation components are first discussed individually and then combined to describe the
total gap variation.

Translational Variation Component:


Translational variation at an assembly gap is characterized by a uniform shift of
the mating part surfaces due to dimensional variation. In order to illustrate this variation
component, a simple flexible part is introduced in Figure 4-5.

43

r r

Mating
Gap Edge


h h

y
z

w w
Figure 4-5 Dimensioned flexible part used for illustration

By looking at how variation within the part tolerances affects the mating gap edge of this
example part, the variation components that contribute to translation can be identified.
These are shown for translation in the y-direction in Figure 4-6, and include dimensional
variation in the part height h and the bend angle .
r

Nominal

Nominal
+

h + h

y
z

y
x

a) Variation in part height

b) Variation in bend angle

Figure 4-6 Illustration of translational variation in flexible part

In this figure, maximum variation is shown at extremes by the shaded part and the
small dashed lines. The nominal part geometry is shown with larger dashed lines. From
Figure 4-6a, it is apparent that variation at the gap edge ( h ) relates directly to
variation in the part height. The variation in the bend angle, in Figure 4-6b, is amplified
44

at the mating edge by the horizontal length r, resulting in a variation of r (for small
angles). The total translational variance of the example part mating edge, in the ydirection, can be written as:
2

2
Ty , part

y
y
= h2 + 2
h

(4-4)

Where the variances of the part height h2 and bend angle 2 can be described in terms
of the part tolerances (assuming the tolerances represent 3 ):
2

h

= ; 2 =
3
3

2
h

(4-5)

The treatment of tolerances as 3 is a common practice when measurement data is not


available, such as in the design phase of product development. If measurement data is
available, the variances in equation 4-5 can be calculated directly from the data. The
sensitivities of variation in the y-direction y to variation in the height h and bend
angle can be written using the mating edge variations shown in Figure 4-6 as:
y
=1
h
y
y = r
=r

y = h

(4-6)

Using the vector-loop methods describe in section 4.2.2, the translational variation
components in equation 4-4 can easily be accounted for using the DLM.

45

Rotational Variation Component:


Twist or rotational variation along a gap is the result of dimensional variation
along a plane that is parallel to the gap. This is illustrated in Figure 4-7, using the
example part from Figure 4-5, with variation in the dimensioned height h along the width
w of the part (z-direction).
w

Nominal
h + h

h h
z

h
y
x
y
z

a) Rotation variation at the gap edge

b) Rotational variation (seen from front)

Figure 4-7 Illustration of rotational variation in flexible part

In section a of this figure the maximum rotation cases, shown by the shaded part and the
small dashed lines, bracket the rotational variation at the gap. This type of variation
would occur if the bend was not made parallel to the bottom edge or if the bottom edge
was not cut straight. A frontal view of the part, in Figure 4-7b, shows how variation of
the height h along the part width results in rotation limits about the x-axis of . By
treating rotation along the z-axis separately from variation in x and y, the 3D analysis is
uncoupled allowing for simplified 2D analysis in the x-y plane. From the part dimensions
in Figure 4-7, the rotational (angular) variation can be calculated by:
2h
= tan 1

(4-7)

46

The rotational variance of the example part mating edge, in the y-direction, can be written
for each fastening point i as:
2

2
Ryi , part

y
= i 2

(4-8)

Again, when measurement data is not available, the angular variance 2 can be
expressed in terms of the variation limit:

=
3

(4-9)

For small angles the sensitivity of variation in the y-direction y i to angular variation
can be expressed using the following equation:
y i = z i

y i
= zi

(4-10)

In equation 4-10, z i represents the distance from point i to the center of the gap line. As
expected, variation in the y-direction due to angular rotation is different at each
fastening node and is dependant upon the nodal position along the mating edge.

Surface Variation Component:


For flexible parts, surface variation represents the surface waviness of compliant
surfaces. Such part variation results from imperfections in tooling surfaces and from
residual stress introduced by forming processes used in part manufacturing. In design,
surface variation is bounded by geometric feature tolerance bands, as illustrated in Figure
4-8.

47

Design
Tolerance:

Resulting
Tolerance Band:

Figure 4-8 Form tolerance limits of surface variation

If surface measurement data is available, the standard deviation at each node can be
calculated directly and used to create a surface variance vector. Otherwise, the tolerance
band can be treated as 3 standard deviations. The variance of the mating surface for any
part P is then simply:

2
SP
= P
6

(4-11)

Where P is the tolerance band along the mating surface of part P. The denominator in
equation 4-11 is 6 instead of 3 because P describes the entire tolerance zone, and not
just half of it.
If measurement data is unavailable, the standard deviation may be estimated using
equation 4-11 for each flexible parts mating gap surface. Surface variation occurring at
other joints in the assembly is treated as form variation that propagates through assembly
joints, and is not included in this step of the analysis. For a uniform tolerance band, the
surface variance is the same at each fastening node i. With n fastening points, the surface
variance vector for each mating part surface P is:

{ } = {1 1 L 1 }
2
SPi

2
SP

(4-12)

48

Determining the surface variance at each fastening point in vector form (for each flexible
gap surface) gives the required surface variation output for step 1 of FASTA. For a nonuniform tolerance band along a flexible surface, equation 4-6 would simply be adjusted to
include the differing variance at each fastening location.

Combining Rigid Variation Components:


Considering the example part in Figure 4-5, the total variance of its mating edge,
in the y-direction, can be obtained by summing the part variation components in
equations 4-4, 4-8 and 4-11:
2
2
y2, part = Ty2 , part + Ry
+ SP
i , part

(4-13)

This combined variance illustrates translational, rotational, and surface variances add
statistically at the mating edge in a single part.
In assembly, variation seen at the gap is a result of statistical accumulation of
variation in multiple parts. Each component of gap variation can be treated for multiple
parts and then combined statistically to represent the total gap variation. For mating parts
A and B, the gap translational variance T2 is obtained by summing the translational part
2
2
variances TA
and TB
, as shown in the following equation:

2
2
T2 = TA
+ TB

(4-14)

The variance of the gap rotational variation R2 can be determined similarly as:
2

=
2
R

2
RA

2
RB


= A + B
3 3

49

(4-15)

In this equation, the part rotational variances are also shown in terms of the angular
variations for parts A and B (see equation 4-9). Rotational variation of parts other than
the mating gap parts propagate through assembly joints, and should be treated separately
as form variation. The total rigid gap variation for mating parts A and B can be written
using equations 4-11, 4-14 and 4-15 as:
2
2
2
Gap
= T2 + R2 + SA
+ SB

(4-16)

It should be noted in this equation that the surface variances for the mating parts are not
combined into a single gap term, because their individual values are needed for
covariance modeling in step 3 of FASTA.
Due to the covariance modeling limitations described earlier, separate variance
vectors for each rigid variation mode are needed as an output from step 1 of FASTA.
Although each variation mode in equation 4-16 can be treated using the DLM, for
conceptual and analysis simplicity only translational variation will be determined using
this method. By treating rotational and surface variations separately, the solution to the

{ }

DLM analysis can be used to create the gap translational variance vector Ti2 . This
process is illustrated using an example assembly in the next section.
The gap rotational variance vector can be determined in radians using equation 415. The angular rotational variation at each mating point i (for n fastening nodes) is
simply:

{ } = {1 1 L 1 } (rad )
2
Ri

2
R

50

(4-17)

In this equation, the subscript n denotes that the vector has n columns. To resolve
angular variation into linear variation, equation 4-10 can be used to express the rotational
variance vector in units of length:

{ } = {z
2
Ri

2
R

z 2 L z n } (length )
T

(4-18)

where z i , again, represents the distance from fastening point i to the center of the gap

{ }

2
line. The surface variation vector SPi
for each flexible mating part P is shown in

equation 4-12.
With the methods presented in this section, each of the required gap variance
vectors for step 1 of FASTA can be determined. The variation modal breakdown also
provides a tool for looking at how assembly variation is truly linked to dimensional
variation.

4.2.4 Example assembly


In order to illustrate the application of the steps of the FASTA method presented
in this and following chapters, the example assembly shown in Figure 4-9 is analyzed. In
this figure the ideal assembly of parts A, B, and C is shown, creating a four sided box
using two formed sheet metal parts and a rigid base. An exaggerated assembly gap is
also shown, resulting from part variation from ideal. The joining of flexible parts A and
B (the closure of the assembly gap) is done using five fastening points along their mating
surfaces.

51

Gap

a) Ideal assembly of parts: A, B, C

b) Gap resulting from part variation

Figure 4-9 Example flexible assembly

The dimensions of each of the parts in the example assembly are shown in Figure
4-10. Each dimension is shown referenced to the corresponding parts Datum Reference
Frame (DRF), shown as a small box with crossed diagonal lines. As noted in Chapter 3,
dimensioning schemes and datum references do play an important role in variation stackup. A vector loop that models assembly interactions between parts must pass through
each parts DRF before continuing on to the next part (see Appendix A).

52

A
B

b
tA

tB

e
j

y
x
z

k
i
Figure 4-10 Example part dimensions

4.2.4.1 Vector loop model of example assembly


In creating a vector loop model of the example assembly, it is assumed that only
forces in the y-direction are used to close the gap. For this reason, only variation in that
direction is considered in the analysis. Also, only variation that contributes to translation
in the y-direction is included in the vector loop since the rotational and gap surface
variations are treated separately. With these assumptions the vector loop model of the
example assembly is shown in Figure 4-11.

53

fi

u1

u2

Gap yi
gi

u4

u3

Figure 4-11 Vector loop model of example assembly

In this figure vectors u1-u4 are not part dimensions seen in Figure 4-10, but instead
represent tooling dimensions and variations that characterize the placement of the
fastening points. These vectors are an important link to assembly tooling and fixtures,
and are required inputs to the FASTA analysis. The kinematic vector loop equation for
the gap at node i in the y-direction is:
G yi = u 2 sin( 180 + + ) + d sin(0 ) + u 3 sin(180) + u 4 sin(180) + k sin(90)
+ a sin(90) + u1 sin( 90 + ) + f i sin(0) + g i sin(0)

(4-19)

4.2.4.2 DLM applied to example assembly


Once kinematic vector loop equations are created, it is possible to take partial
derivatives of these equations with respect to the independent variables. This is generally
easy to do, since there are only sine and cosine terms to differentiate. These partial
derivatives can then be used to populate the DLM matrices for analysis. For the example
assembly, the partial derivatives of equation 4-4 are:

54

G yi
u 2
G yi
d
G yi
u 3
G yi
u 4
G yi
k
G yi
a

= sin(180 + 2 )

G yi
u1
G yi

= sin(0 )
= sin(180) = 0
= sin(180) = 0
= sin(90) = 1

G yi

G yi
f i
G yi
g i

= sin(90 + )
= u1 cos(90 + )
= 2u 2 cos(180 + 2 ) d cos(0 )

(4-20)

= sin(0) = 0
= sin(0) = 0

= sin(90) = 1

Each of these partial derivatives is evaluated using nominal dimensional values.


Generally these values can be obtained from CAD models or drawings of the parts and
tooling being analyzed. For the example assembly, the nominal dimensions and
variations about the nominal are listed for the tooling/fixtures in Table 4-1 and for parts
A, B and C in Table 4-2.
Table 4-1Tooling/fixture dimensions for example assembly
*Variable Vector Lengths (in)

Dimension

Nominal

Variation

units

u1

9.00

0.15

in

Vector

Nominal

Vector

Nominal

u2

1.00

0.15

in

f1

g1

u3

1.00

0.15

in

f2

g2

u4

9.00

0.15

in

f3

g3

fi

0.15

in

f4

10

g4

10

gi

0.15

in

f5

13

g5

13

55

Table 4-2 Part dimensions for example assembly


Part

Dimension
a
b
c

Nominal
10
10
14

Variation
0.1
0.1
0.3

units
in
in
in

tA

d
e
h

0.025
90
0
9
1.5
14

0.003
5
0.25**
0.1
0.1
0.3

in
deg
in
in
in
in

0.003
5
0.15**
0.2
0.3
0.08

in
deg
in
in
in
in

tB
0.04
90

i
10
C
j
14
k
1
** Variation = (tolerance zone)/2

Using the values contained in these tables, and the partial derivatives in equation
4-20, it is now possible to create the DLM vectors and matrices needed in this analysis.
Again, there will be no [D] matrix because there are no closed loops required to model
the variation in the example assembly, and no [F ] matrix because the form variation at the
gap surface is being treated separately. The vector of manufacturing variations, {X } , can
be evaluated by simply inserting the known variation values. For the example assembly
this is:

56

u 2 0.15
d 0.1

u 3 0.15

u 4 0.15
k 0.08

{X } = a = 0.1
u 0.15
1

f i 0.15

g i 0.15
0.0873

0.0873

(4-21)

Of special note in equation 4-21 are the angular variations and , which must be in
radians in order to maintain the correct dimensions during matrix algebra. Matrix [C] for
the example assembly is mathematically represented as:
G y1

u 2
G y 2
u
2
G
[C] = y 3
u
G 2
y4
u 2
G
y5
u 2

G y1

G y1

G y1

G y1

G y1

G y1

G y1

G y1

G y1

d
G y 2

u 3
G y 2

u 4
G y 2

k
G y 2

a
G y 2

u1
G y 2

f i
G y 2

g i
G y 2

G y 2

d
G y 3

u 3
G y 3

u 4
G y 3

k
G y 3

a
G y 3

u1
G y 3

f i
G y 3

g i
G y 3

G y 3

d
G y 4

u 3
G y 4

u 4
G y 4

k
G y 4

a
G y 4

u1
G y 4

f i
G y 4

g i
G y 4

G y 4

d
G y 5

u 3
G y 5

u 4
G y 5

k
G y 5

a
G y 5

u1
G y 5

f i
G y 5

g i
G y 5

G y 5

u 3

u 4

u1

f i

g i

G y1


G y 2

G y 3

G y 4

G y 5


(4-22)

Solving for the partial derivatives at the nominal dimensions this becomes:
0
0

[C] = 0

0
0

1
1
1
1
1

0
0
0
0
0

0
0
0
0
0

1
1
1
1
1

1
1
1
1
1

0
0
0
0
0

57

0
0
0
0
0

0
0
0
0
0

9
9
9
9
9

2
2
2

2
2

(4-23)

In comparing equations 4-22 and 4-23, there are many things that can be
observed. First, it is apparent that all of the rows are identical. This makes sense since
the only things that could be different from row to row are the partial derivatives with
respect to the variable vector terms (fi, gi). These, however, are zero since they are
perpendicular to the gap coordinate direction of interest, leaving each row to be the same.
Values of 1 in the matrix suggest that the corresponding dimensional variation lies
along the y-direction and thus contributes directly to the gap clearance. Values of zero
suggest that the related variation does not contribute to variation in the y-direction at all.
The last two columns in the matrix are large, corresponding to angular variations and the
length upon which they act.
Without any close loops and with no rigid form variation components, equation 42 simplifies to:
{G yi } = [C]{X}

(4-24)

The tolerance sensitivity matrix for the system is then simply[S d ] = [C] . Using this
sensitivity with the dimensional and form tolerances in equation 4-21, the gap variation in
the y-direction can be calculated. This is done by applying a statistical variation stack-up
model as shown in equation 3-19. Leaving the tolerances as variables, this equation is:
1

(0 u 2 )2 + ( 1 d )2 + (0 u 3 )2 + (0 u 4 )2 2

2
2
2
2
G yi = + ( 1 k ) + (1 a ) + (0 u1 ) + (0 f i ) Tasm yi

2
2
2
+ (0 g i ) + (9.25 ) + (1.5 )

(4-25)

Assuming that the assembly tolerance represents six standard deviations (3), then the
variance is calculated by:
58

yi

Tasm yi
=
3

(4-26)

Substituting in the variation values, the solutions to equation 4-25 and 4-26 are:
Tasm yi = 0.6737

(4-27)

yi = 0.0504

As shown in Figure 4-1 at the beginning of the chapter, the desired outputs of this
step of FASTA are the mean and variation of the gap. With gap rotational and surface
components excluded from the vector loop model, the variance solution in equation 4-27
is simply the translational variation seen at the gap. The vector of translational variance
at each fastening point along the gap is then:

{ } = { } = {0.0504
2
Ti

2
yi

0.0504 0.0504 0.0504 0.0504}

(4-28)

The mean gap solution for this example is much easier to obtain than the variance
solution. Since the assembly is described by an open vector loop, evaluating the vector
loop equation at nominal dimensions provides the mean solution. Using the nominal
values in Table 4-1 and Table 4-2 in equation 4-19, the mean separation in the y-direction
at each node is:

{ i } = {0

0 0 0 0}

(4-29)

A zero mean solution suggests that the parts are designed to fit together nominally,
though manufacturing variation cause deviations from this ideal described by the nonzero results of equation 4-28.

59

4.2.4.3 Rotational and surface variation in example assembly


Rotational Variation:
Rotational variation at the gap in the example assembly is a result of variation of
parts A and B parallel to the gap line. Since only variation in the y-direction is being
analyzed, only rotation about the x-axis needs to be looked at. From Figure 4-10 it is
apparent that variation in dimensions a and d contribute to rotational variation at the gap
for parts A and B respectively. Using equation 4-7 and the part dimensions, the
rotational variations of parts A and B are:
2a
A = tan 1
= 0.0143 rad
c

(4-30)

2d
B = tan 1
= 0.0143 rad
h

(4-31)

and,

The variance of gap rotational variation can then be solved using equation 4-15:
2


2
= A + B = 4.535 10 5 (rad )
3 3
2
R

(4-32)

It should be noted that equations 4-30 to 3-22 have angular units of radians, and should
not be evaluated in degrees. Applying the result of equation 4-32 in equation 4-17, the
vector of rotational variation (with 5 fastening points) is:

{ } = 4.535 10 {1 1 1 1 1} (rad)
2
Ri

60

(4-33)

Surface Variation:
As discussed earlier, surface variation is bounded by the form tolerance zone
applied to each mating part surface. Since both of the mating gap surfaces are on flexible
parts (A and B), a surface variation vector must be created for each. In Figure 4-10,
geometric feature tolerances A and B are applied to respective part surfaces. The
variance for the mating surfaces of parts A and B can be calculated using equation 4-11
as:
2

2
SA


= A = 0.00694
6

2
SB


= B = 0.0025
6

(4-34)

and,
2

(4-35)

From equations 4-25, 4-24 and 4-8, the surface variance vectors for parts A and B are:

{ } = 0.00694{1 1 1 1 1}
{ } = 0.0025{1 1 1 1 1}

2
SAi
2
SBi

(4-36)

The y-dimensional variation of the example assembly gap is now completely


described by equations 4-28, 4-33, and 4-36. Though the variation has been split into
separate components for use in later steps of FASTA, the total gap variance is simply the
sum of the translational, rotational and surface variances. If such a summation was
desired, the rotational variation would first need to be converted from angular to
translational units using equation 4-18.

61

4.3 Other Considerations for Flexible Assemblies


In analyzing flexible assemblies, there are several special cases and
considerations that warrant discussion. Topics included in this section are: positional
variation of nonparallel surfaces, tooling contribution, multiple gap variation propagation,
and obtaining geometry and variation data. It is hoped that by treating these topics a
better understanding of the intricacies of modeling flexible assembly variation will be
gained and that the path for future work will be paved.

4.3.1 Positional variation of nonparallel surfaces


In treating the modeling of gap variation along flexible surfaces, there exist
limited cases when positional variation of the fastening points could contribute to
variation along multiple coordinate axes. Such an instance occurs when mating surfaces
are not nominally parallel to each other. Take for example the simple case of mating
surfaces A and B shown in Figure 4-12.
dfi

dxi

fi

dyi

A
B

y
x

gi

Figure 4-12 Nonparallel surface variation

Here, a positional variation dfi of fastening location i causes variation components dxi and
dyi. This means that G yi f i 0 and variation of vector fi contributes the gap variation
at node i.
62

The example of nonparallel surfaces presented in Figure 4-12 is easily taken into
account using variable vectors fi and gi in the DLM (see section 4.2.2) provided that the
orientation of these vectors is known. If either the slope or angle of a vector at node i is
known, then the variation component can be calculated. For example, to solve for
G yi f i given the slope at i (mi), angle can be solved for using trigonometric
equations:
dy
= mi
dx
= tan 1 (mi )
tan( ) =

(4-37)

The variation in y is then calculated:


sin( ) =

dy
df i

(4-38)

dy = df i sin( ) = dG yi
From this equation it is apparent that the variation contribution of dfi in the y-direction is
accounted for in the matrix algebra of [C]{dX } where the variation component in matrix

[C] becomes:
G yi
f i

= sin( )

(4-39)

and dfi is contained in vector {X } .


This technique can be extended to nonparallel curved surfaces by evaluating the
derivative of the curve equation at each fastening point to approximate the slope for small
variations. If a curve equation is not available, polynomial or other curve fitting methods
could be used to approximate the slope if deemed significant for the analysis.
63

4.3.2 Tooling contribution


Due to the compliant nature of the parts upon which variation analysis is being
applied, it is important to consider the influence of fixtures and tooling on assembly. In
the example assembly vector loop presented in Figure 4-11, recall that the dimensional
values for vectors u1-u4 were determined by the fixtures and tooling used to locate the
fastening points between mating parts. Since these are independent of the part
dimensions, meaning that they cannot be calculated simply by knowing the dimensions of
the parts, they are required inputs to the analysis and must be known. Since fixtures are
usually designed specifically to position parts accurately for fastening and joining, the
fastening locations (and variation limits) can often be pulled directly from the design
drawings. In the case where production tooling is in use, it is also possible to use
accurate measuring devices, such as a Coordinate Measuring Machine (CMM) to obtain
the fastening positions directly from the fixtures. The methods used to obtain these
dimensional values will depend upon the availability of information and the desired
accuracy of the analysis.
In addition to the positioning effects of tooling on the assembly process, there are
also possible secondary effects from tooling that may need to be considered. For
example, Figure 4-13 shows two cantilever beams deflecting under an applied tooling
force F. In addition to bending deformation, note that there is also a separation of the
two beam ends labeled as a distance m.

64

Upper Beam

Resultant forces if
fastened & released:

F
T
N

Lower Beam

Figure 4-13 Deformation of cantilever beams [Liu 1995]

Fastening the two beams together and then removing force F (equivalent to removing
parts from the fixture), would result in the spring-back force components at the fastening
point as shown in the right side of the figure. These forces at the fastening point would
include not only a force normal to the surface (N), but also a tangential shearing force (T)
and a moment (M). The FASTA method presented in this thesis can only model either
the normal force or the tangential force by itself, making the assumption that all other
forces are negligible. If needed, each force could be calculated separately and then
combined using superposition. A detailed derivation of this process is left to future work.

4.3.3 Multiple gap variation propagation


When modeling compliant parts in an assembly, it quickly becomes apparent that
most assemblies contain more than one flexible joint. For example, the assembly
presented in Figure 4-9 has flexible surfaces involved at all three joints between parts.
The question then arises: How does the closing and fastening of one flexible joint affect
another? Or stated in another way: Does gap variation propagate rigidly or elastically
through multiple flexible joints? These questions regarding flexible joints are put
forward again and discussed in detail in Chapter 7. However, it is worthwhile to discuss
some issues relating to multiple gaps at this point.
65

A familiar example of closing multiple flexible gaps is the assembly of a


trampoline. If springs are placed on one side of a square trampoline, then the first spring
put on the opposite side will experience the greatest load and consequently the greatest
probability of yielding. This is graphically shown in Figure 4-14a.

a) Uneven spring assembly

b) Even spring assembly

Figure 4-14 Trampoline order of assembly example

If instead, the springs are placed evenly on opposite sides during assembly (Figure
4-14b), then much less force is required to connect each springs to the trampoline and
yielding is not likely. Provided that no yielding occurs in the springs, both examples in
Figure 4-14 will result in an equilibrium assembly as shown in Figure 4-15.

Figure 4-15 Equilibrium trampoline assembly

66

From this illustration it is apparent that the path to final assembly may be more
critical than the final equilibrium solution. If plastic deformation or yielding of
compliant parts is an issue during assembly, the order of assembly should be reviewed.
A method for analyzing the effects of assembly order is left to future work.

4.3.4 Obtaining geometry and variation data


As required inputs to this and other steps of FASTA, geometry and variation data
are keys to the success of the analysis. However, obtaining this information is not always
an easy task. A brief discussion on possible challenges to obtaining geometry and
variation values is presented here to prepare the reader in advance for potential obstacles.
Although some form of assembly design data is often available, it is not always in
the desired format. Design drawings may be difficult to interpret or may contain
unconventional dimensioning schemes that are challenging to model using vector loops.
This is particularly true when the only available drawings predate modern design
practices and techniques. CAD models of parts to be analyzed can be very useful,
especially in creating finite element models (see Chapter 5). However, data exchange
between graphics programs is prone to difficulties even with standard formats like IGES.
CAD data locked in outdated program revisions is also a point of concern even when
dealing with parts modeled just a few years ago.
Even if part geometry and variation data do exist, whether in the form of design
drawings or measurement statistics, simply getting access to them can be difficult.
Boundaries between design and manufacturing disciplines may present obstacles in data
acquisition. Also, it is possible that people that dont understand the need or implication
67

of dimension and variation data for analysis may prevent or delay data flow within a
company or organization. Whatever challenges there may be, obtaining useful data for
analysis is a necessary step in improving product quality and should be treated as such.

4.4 Summary
In this chapter tools have been introduced to statistically determine gap mean and
variation at each fastening point of a flexible joint. This was done by adapting the DLM
method using variable length vectors along mating flexible surfaces. The gap variation
was also separated into translational, rotational and surface variation components. An
example assembly has been introduced and analyzed to illustrate the process of the
analysis. Other considerations in modeling the assembly of compliant parts have also
been addressed. Table 4-3 summarizes the inputs, outputs and analysis steps required to
determine misalignment. A process flow diagram of these steps is also presented in
Figure 4-16. This is equivalent to Figure 4-1, but now includes much more detail.

68

Table 4-3 Summary of steps to determine misalignment

Inputs: part geometry, part variation


2
Outputs: gap variation components: Ti2 , Ri2 , SPi
; and gap mean: { i }
Analysis Steps:

{ }{ }{ }

A. Variation Component Solutions


Translational Variation
1. Create vector loop model
2. Write kinematic vector loop equations
3. Take partial derivatives of vector loop equations
4. Populate DLM matrices and vectors
5. Apply Statistical stack-up model to open loop equations
6. Calculate variance at each fastening point
7. Create translational variance vector
Rotational Variation
1. Calculate rotational variation of mating gap parts
2. Statistically combine part rotational variations
3. Create rotational variance vector
Surface Variation
1. Identify tolerance bands at each flexible gap surface
2. Calculate surface variance for each flexible part P
3. Create a surface variance vector for each flexible part
B. Mean Gap Solution
1. Solve for mean gap value at each fastening point using:
a. Vector loop equation and nominal dimensions
b. CAD assembly model
2. Output mean gap vector

69

FASTA Step 1: Determine Misalignment


Part Variation

Part Geometry

Inputs

Outputs

VL Model
Mean

VL Equations

Solve for:
i

Partial
Derivatives

CAD

DLM Matrices
& Vectors

STA

{i }

Translational Variation

Statistical
Stack-up Model

Calculate:
Ti2

{ }
2
Ti

Rotational Variation

Determine
for gap parts

Solve for:
R of gap

{ }
2
Ri

Surface Variation

Determine P
for each flexible
gap surface

Calculate:
2
SPi

{ }
2
SPi

Figure 4-16 Process flow diagram for statistically determining misalignment

70

CHAPTER 5

Finite Element Modeling


Of Compliant Parts

5.1 Introduction
The second step of the FASTA process is to model flexible part compliance using
finite element techniques. Referring back to the general force-deflection equation shown
at the beginning of Chapter 4 (equation 4-1), this analysis step requires the calculation of
each compliant parts stiffness matrix [K]. For simple geometry it is possible to calculate
a components stiffness matrix by hand. However, parts containing multiple bends, cuts
and curvature require more complicated techniques. Finite element analysis provides the
tools needed to model and analyze even complex parts with relative ease. Using
commercially available FEA software, it is possible to obtain a representative stiffness
matrix for any elastic part with known geometry.
In this thesis ANSYS, a widely available FEA software package, is used to
illustrate the application of the analysis steps presented. While the general discussion and
processes shown in this chapter can be implemented using other FEA packages, detailed
instruction for analysis will only be given for ANSYS.

71

In order to adequately model a parts compliance, both part geometry and material
properties must be known. An overview of step 2 is shown in Figure 5-1, including both
the needed inputs and desired output.

FASTA Step 2 Overview


Part Geometry

Material
Properties

Inputs

2. FEM

[Keq]

Output

Figure 5-1 Step 2 process summary

This chapter begins with a discussion of assumptions used in the analysis. Background
is presented and the steps needed to model flexible parts using FEA are illustrated using
the example assembly from Chapter 4.

5.2 Analysis Assumptions


As is always the case, this analysis uses some simplifying assumptions in the
modeling of part compliance. Since nonlinear and plastic deformations result in a
variable stiffness matrix, all deformations will be assumed to be both linear and elastic.
These are good assumptions given that tolerance analysis usually involves small relative
displacements and resulting deformations. Based upon these assumptions, the stiffness
72

matrix [K] is constant over the analysis range and can thus be evaluated using nominal
geometry and then applied to calculate deformation forces.
It is also assumed in this chapter that the reader has a general understanding of the
terms and basic use of finite element analysis. In order to apply the processes presented,
access to FEA software, such as ANSYS, is required.

5.3 Background
A stiffness matrix, sometimes called an elasticity matrix, mathematically
represents the elastic material properties and interactions within its representative part.
Quite literally it describes how stiff a part is by relating force to displacement. The
simplest example of stiffness can be seen by looking at a simple bar in tension. If a
tensile force F is applied to a bar with an initial length L and cross-sectional area A, the
resulting elongation is given by as shown in Figure 5-2.

L
A
F

Figure 5-2 Example of stiffness of bar in tension

Using the definitions of and relations between stress and strain, stress can be written as:
= E = E

(5-1)

73

In this equation E represents the modulus of elasticity of the beam. Since stress is equal
to force over area (for tensile loading), the results of equation 5-1 can be applied to get:
F = A =

E
A
L

(5-2)

By simply rearranging equation 5-2 it can be easily seen that the equation relates force to
displacement with stiffness as given in equation 5-3.
K=

EA
L

(5-3)

From this simple example it is apparent that the stiffness K is dependant upon both
material properties (E) and part geometry (L and A).
In the context of the FASTA analysis, compliant parts are often bent into contact
with other parts. One of the simplest examples of stiffness in bending is that of a beam
with both applied bending forces and moments at each end as illustrated in Figure 5-3.

F1 , u1

F2 , u2

M1 , 1

M2 , 2
Figure 5-3 Example of stiffness for 2-D bending

In this case the relationship between force and displacement is given as:
6
L2
F1
3
M

1 2 EI L
=
L 6
F2
L2
M 1
3

3
L
2
3
L
1

6
L2
3
L
6
L2
3
L
74

3
L u
1
1
1
3 u2

L
2
2

(5-4)

The variable I in equation 5-4 represents the moment of inertia for the beam. With 4
possible ways of displacing the beam (u1, 1, u2, and 2) a 4x4 matrix is needed to
describe its stiffness. This correlation between the degrees of freedom (DOF) and the
matrix size holds true for any case. This can become a problem when modeling a
complex part with thousands of DOF, a challenge addressed by the next section. It
should also be noted that the stiffness matrix is always square and symmetric.

5.3.1 Matrix reduction


As shown in the examples of the last section, as the number of degrees of freedom
increases, so does the size of the stiffness matrix. This is particularly challenging when
using finite element models of parts, since each node can have up to 6 degrees of freedom
(translation in and rotation about the x, y, and z coordinate axis). As the number of
elements and nodes in a model is increased to improve FEA accuracy, the corresponding
stiffness matrix is also enlarged. The result is increased computational time when
applying the stiffness matrix to calculate the solution.
Matrix reduction, a matrix partitioning technique, provides a tool for significantly
reducing a matrix size without loss of accuracy. This is done by reducing the matrix to
include only the nodes of interest, while still accounting for the influence of the other
nodes in the part. Such a reduced matrix is often called a super-element matrix. The
derivation for matrix reduction can be found in [Crisfield 1986], and is summarized in
this section.
In order to reduce the matrix, one must first determine which nodes are important
for the analysis. These are selected as boundary nodes and generally include fastening
75

points, locations of applied forces and nodes with applied constraints. All other nodes are
referred to as interior nodes and are important only because of their influence on the
boundary nodes. Figure 5-4 shows a simple example of the boundary and interior nodes
for a plate in bending.
Nodal Constraints

boundary nodes

interior nodes
Figure 5-4 Boundary and interior nodes of a plate in bending [Tonks 2002]

With boundary nodes selected, the matrix form of the spring equation can be partitioned
such that the boundary forces (Fb) and displacements (b) are separated from the interior
forces (Fi) and displacements (i). The force-deflection equation can then be written as:
Fb K bb
=
Fi K ib

K bi b

K ii i

(5-5)

where the stiffness matrix has been divided into boundary node stiffness (Kbb), interior
node stiffness (Kii), and stiffness interactions between the boundary and interior nodes
(Kbi and Kib). Equation 5-5 can be written as two linear equations:

{Fb } = [K bb ]{ b } + [K bi ]{ i }

(5-6)

{Fi } = [K ib ]{ b } + [K ii ]{ i } = 0

(5-7)

76

Note that since there are no external forces on the interior nodes, equation 5-7 is equal to
zero. Solving this equation for the vector of internal displacements gives the following
equation:

{ i } = [K ii ]1 [K ib ]{ b }

(5-8)

Substituting equation 5-8 into equation 5-6, the force vector at the boundary nodes is:

{Fb } = [K bb ]{ b } [K bi ][K ii ]1 [K ib ]{ b }
1
= ([K bb ] [K bi ][K ii ] [K ib ]){ b }

(5-9)

From equation 5-9 it is apparent that the reduced stiffness matrix can be written as:

[K red ] = [K bb ] [K bi ][K ii ]1 [K ib ]

(5-10)

This reduced stiffness matrix (or super-element matrix) contains the influence of the
interior nodes, but is reduced to the size of the boundary node stiffness.
For FASTA, a super-element matrix needs to be created for each compliant part.
With multiple parts, matrix partitioning and reduction can be time consuming, especially
if done by hand. ANSYS automates this task using sub-structuring analysis to output a
parts super-element matrix. This automated process is introduced later in the chapter,
and described in detail in Appendix C.

5.3.2 Equivalent stiffness


When modeling the joining of two compliant parts, it is possible to create an
equivalent stiffness matrix that represents the stiffness of the combined parts at
equilibrium. Creating such a matrix allows for the calculation of the forces seen by each
part at equilibrium position, as well as the corresponding part displacements. It can also
77

be useful when dealing with complex systems of multiple flexible parts to simplify the
model.
The joining of two springs (A and B) is shown in Figure 5-5 to illustrate how
equivalent stiffness can be calculated.

KA

FA

KB
FB

Equilibrium: FA = FB
Gap: 0 = A B
Spring Equations: FA = K A A
FB = K B B
Figure 5-5 Spring example of equivalent stiffness [Merkley 1998]

Combining the equilibrium, gap and stiffness equations shown in this figure, the
equilibrium displacements (A and B) can be written in terms of the stiffness values (KA,
KB) and the overall gap (0). In matrix form, these equations are:

{ A } = [([K A ] + [K B ])1 [K B ]]{ 0 } = [K RA ]{ 0 }

(5-11)

{ B } = [([K A ] + [K B ])1 [K A ]]{ 0 } = [K RB ]{ 0 }

(5-12)

Solving for the equilibrium force using either equation 5-5 or 5-6 and the corresponding
stiffness equation, the matrix form of the solution is:

{FA } = {FB } = [[K A ]([K A ] + [K B ])1 [K B ]]{ 0 } = [K eq ]{ 0 }

78

(5-13)

From this equation it is apparent that the equivalent stiffness matrix is calculated in terms
of [K A ] and [K B ] using a series of matrix operations. Separating it from equation 5-13,
[K eq ] can be written as:

[K ] = [K ]([K ] + [K ]) [K ]
1

eq

(5-14)

If one of the parts is rigid (essentially infinitely stiff), then equation 5-14 can be reduced
to show that the equivalent stiffness is equal to the stiffness of the flexible part.
With the equations presented in this section, gap distances between compliant
parts can be used along with part stiffness matrices to calculate equilibrium
displacements and forces. As discussed at the beginning of this chapter, the desired
output of this step in the FASTA process is[K eq ] . In order to solve for this matrix, the
part stiffness matrices must first be calculated.

5.4 FEA Modeling of Flexible Parts


Finite element analysis greatly simplifies the calculation of part stiffness matrices
by using linearizing assumptions across finite portions of the surface. This section is a
general overview of the process of modeling flexible parts using FEA. The intent of the
FEA model at this point in the analysis is to obtain the reduced stiffness matrix for each
compliant part. A summary of the steps required to do so is presented in Table 5-1.
Each of these steps is discussed and applied to the example assembly from Chapter 4
using ANSYS. Modeling issues and suggestions for FEA are included in Appendix B
for the interested reader.

79

Table 5-1 Summary of FEA modeling steps

Steps for FEA Modeling of Flexible Parts


1. Define nominal geometry
2.

Input material properties

3.

Select Element type and properties

4.

Mesh the model

5.

Output the reduced stiffness matrix

5.4.1 Define nominal geometry


The first task in modeling flexible parts is to define the nominal geometry. In
order to do this, the dimensions and thickness for each component must be known.
Models can be made directly using part drawings and data, or CAD models can be
imported into the FEA software using IGES or other accepted formats. Care must be
taken that imported files maintain their original geometry and are free of errors resulting
from improper cross-package translation. When familiar with the FEA softwares user
interface, this step should not require much time.
Using the nominal dimensions defined in Table 4-2, parts A and B of the example
assembly can be modeled easily in ANSYS. This was done by creating keypoints and
connecting them with straight lines to define the profile. The lines were then extruded in
the negative z-direction to create part areas. Thickness is added later in the process. The
resulting images for both parts are shown in Figure 5-6.

80

a) Nominal model of part A

b) Nominal model of part B

Figure 5-6 Nominal model of flexible parts in example assembly (created in ANSYS)

5.4.2 Material properties


Since FEA software packages model material interactions, material properties are
required inputs. The number and type of inputs vary based upon the complexity of the
material used, ranging from simple isotropic metals to complex laminates with
orthotropic characteristics. These values can usually be found in material property tables,
on property sheets obtained from material producers, or if need be from testing.
Commonly used (or assumed), linear isotropic materials require only the modulus of
elasticity (E) and Poissons ratio () to describe the properties of the material. The
material properties for parts A and B of the example assembly are shown in Table 5-2.
Table 5-2 Material properties of parts A & B

Part A
Part B
Material: Carbon steel Carbon steel
30.0 Mpsi
30.0 Mpsi
E:
0.28
0.28
:
81

5.4.3 Element selection


FEA software uses element models to approximate the behavior of finite sections
of each part being modeled. In order to proceed, it is necessary to pick from a list of onehundred or more different elements the one that best fits the characteristics of the parts
being modeled and the analysis being done. For the purpose of determining the stiffness
and reaction to loading of compliant parts, the possible elements to be used are limited to
those specifically designed for structural analysis. 3D structural element types available
in ANSYS include points, lines, beams, pipes, shells and solids. A single example of
each of these is shown in Figure 5-7. From the selection of structural elements, shell
elements are a good choice for modeling thin flat partsexactly the type of parts
analyzed using FASTA.
Structural Point

Structural 3-D Line

Structural 3-D Beam

Structural Mass

Spar

Elastic Beam

MASS21

LINK8

BEAM4

1 node 3-D space


DOF: UX, UY, UZ
ROTX, ROTY, ROTZ

Structural Pipe

2 nodes 3-D space


DOF: UX, UY, UZ

Structural 3-D Shell

2 nodes 3-D space


DOF: UX, UY, UZ
ROTX, ROTY, ROTZ

Structural 3-D Solid

Elastic Straight Pipe

Shear/Twist Panel

Structural Solid

PIPE16

SHELL28

SOLID45

2 nodes 3-D space


DOF: UX, UY, UZ
ROTX, ROTY, ROTZ

4 nodes 3-D space


DOF: UX, UY, UZ
ROTX, ROTY, ROTZ

8 nodes 3-D space


DOF: UX, UY, UZ

Figure 5-7 Examples of 3-D structural elements in ANSYS

82

5.4.3.1 Shell elements


Shell elements are specifically designed to model the characteristics of thin sheets
of material that may be manufactured (pressed, bent, rolled, etc.) in any form. Specific
selection of a shell element is determined based upon the geometry and material
characteristics being modeled. Figure 5-8 shows three suggested shell elements available
in ANSYS for use in modeling flexible parts, each with distinctive attributes.
Elastic Shell

Structural Shell

Linear Layered
Structural Shell

SHELL63

SHELL93

SHELL99

4 nodes 3-D space


DOF: UX, UY, UZ
ROTX, ROTY, ROTZ

8 nodes 3-D space


DOF: UX, UY, UZ
ROTX, ROTY, ROTZ

8 nodes 3-D space


DOF: UX, UY, UZ
ROTX, ROTY, ROTZ

Figure 5-8 Suggested Shell Elements for flexible part modeling

As shown in Figure 5-8, SHELL63 has only 4 nodes allowing for minimal
computation time at the expense of some accuracy. It is a planar element, but can
approximate curved surfaces well provided that each flat element does not cover more
than 15 of an arc. For more aggressively curved parts, SHELL93 provides an accurate
alternative with no limitations on angle. It also has 8 nodes, allowing for more accurate
modeling at the cost of computational time. Finally, SHELL 99 allows for the modeling
of layered materials, such as those seen in metal and composite laminates. More detail
on the assumptions and inputs for these element and others can be found in ANSYS
documentation. Since example parts A and B both consist of planar sections (Figure 5-6)
and are not laminates, SHELL63 appears to be a good choice for to modeling each part.

83

Once a shell element has been selected in ANSYS, the thickness of the element
(or part thickness) is then input as a Real Constant. While it is possible to assign a
different thickness at each node, generally the material thickness is constant throughout
the element. From Table 4-2 the thicknesses of the example parts can be obtained and
input into the model.

5.4.4 Mesh part


With the element type selected and the material properties and constants set, the
next step is to mesh the parts. Meshing involves segmenting the nominal geometry with
an arrangement of nodes and elements. In the context of the analysis to be done, it is
important that nodes are place at each of the fastening points. This can be done by
careful sizing and placement of elements in the mesh, or by using hardpoints to constrain
the node placement.
Automated meshing is a common feature of many FEA software packages,
allowing for simplified analysis setup. ANSYS provides two automated meshing
options: mapped meshing and free meshing. Mapped meshing is useful for simple
geometry, like rectangles, where identical rows of elements can be laid out across the
surface of the part. Free meshing is not as ordered, allowing each element to adjust in
shape to match complex geometry, such as seen with holes and chamfers. Both methods
can be combined to mesh a complex part on different segments provided that the nodes at
each segment boundary coincide. Element size for automated meshing is controlled by
either specifying desired element sizes or by specifying the number of elements to be
used along each edge of a part.
84

For analysis accuracy, it is desirable that each element is as close to square as


possible in order to best match the linearizing assumptions used. Mesh size also plays a
significant role in analysis accuracy, with smaller elements providing more step
approximations to model real part behavior. Smaller elements, however, result in more
nodes and more computational time. Meshing a part model for both accurate and
streamlined analysis is an art in and of itself, learned with practice and experience.
Since parts A and B consist of rectangular segments, they can be easily map
meshed in ANSYS. With an element edge-length of 0.5 inches, the meshed parts are
shown in Figure 5-9.

a) Meshed model of part A

b) Meshed model of part B

Figure 5-9 Flexible parts in example assembly map meshed with a 0.5 inch edge-length

5.4.5 Reduced stiffness matrix output


With nodes placed at the fastening points of each flexible part, the next step is to
obtain the stiffness matrix for each part and reduce it. For FEA software that outputs a
85

full stiffness matrix it is necessary to partition and reduce each matrix as described in
section 5.3.1. ANSYS greatly simplifies this step by automating matrix reduction using
sub-structuring analysis. The super-element matrices for mating compliant parts are then
combined to create an equivalent stiffness matrix [Keq]. In this section sub-structuring in
ANSYS is briefly introduced and applied to example parts A and B. An equivalent
stiffness matrix for these parts is also calculated.
In ANSYS sub-structuring, constraints are first applied and boundary nodes
(called master degrees of freedom) are selected on each meshed part. With the selection
of boundary nodes, ANSYS also requests the degrees of freedom to be considered at
each node, allowing for further matrix reduction. For compliant parts A and B,
constraints (for all DOF) were placed where they mated to the rigid base (part C) and
boundary nodes were placed where they ideally join together with DOF only in the ydirection. These are shown labeled in Figure 5-10.

a) Part A

b) Part B
Figure 5-10 Master DOF and constraints applied to parts A and B

86

With constraints and master degrees of freedom set, a sub-structuring analysis in


ANSYS outputs the reduced stiffness matrix for each part, along with other information,
in a text file. The output file is not ideally formatted for use, but can be parsed by hand
or using a simple C-program to obtain just the matrix values for later use. It should be
noted that a certain amount of bookkeeping is required to keep track of node numbers and
locations within the output file.
Having only 5 degrees of freedom each, one at each master DOF, the reduced
stiffness matrices for parts A and B can be obtained from ANSYS and are both simply
5x5. These matrices are shown graphically in Figure 5-11. While they appear to be
similar in this figure, it is apparent from the vertical scale that the stiffness of part B is
about ten times that of part A. This is due to the difference in thickness between the
parts, the stiffer one being the thicker of the two.

a) Part A

b) Part B
Figure 5-11 Reduced stiffness matrices for parts A and B (scale in pounds/inch)

Using equation 5-14, the reduced stiffness matrices for parts A and B can be
combined to form an equivalent stiffness matrix. The result is shown graphically in
87

Figure 5-12. Comparing the equivalent stiffness matrix [Keq] with the reduced stiffness
matrices [K red A ] & [K red B ] , it is apparent that its magnitude is slightly lower than that of
part A. This makes sense since [Keq] represents the relation between force and
displacement for the combined parts.

Figure 5-12 Equivalent stiffness matrix (scale in pounds/inch)

In this section a brief introduction and example of sub-structuring in ANSYS has


been given. A more detailed guide on this subject is presented in Appendix C.

5.5 Summary
In this chapter background on stiffness matrix reduction and equivalent stiffness
has been presented. The process of using FEA to obtain reduced stiffness matrices was
enumerated and illustrated using flexible parts from the example assembly in ANSYS.
With this process, nominal geometry and material properties are used to model part
compliance. The output to this step of FASTA is an equivalent stiffness matrix for
mating flexible parts, or the reduced stiffness matrix for a flexible part to be joined to a

88

rigid component. The process is summarized in Table 5-3, and illustrated as a flow
diagram in Figure 5-13.
Table 5-3 Summary of steps to FEM compliant parts

Inputs: part geometry, material properties


Output: [Keq]
Analysis Steps:
A. Model each compliant part with FEM
1. Define nominal geometry
2. Input material properties
3. Select element type and properties
4. Mesh model with nodes at fastening points
5. Solve for stiffness matrix
B. Reduce stiffness matrices
6. Select boundary nodes and degrees of freedom to include
7. Calculate reduced stiffness matrix
C. Solve for equivalent stiffness at assembly gaps [Keq]

89

FASTA Step 2: FEM of Compliant Parts


Part Geometry

Material Properties

Inputs

Nominal Model
Mesh

FEM

Solve for:
Stiffness

[K]

Select Boundary
Nodes, DOF

Repeat for each


compliant part

Reduce
Matrix

[Kred]

Solve for:
Equivalent
Stiffness

[Keq]

Output

Figure 5-13 Process flow diagram for FEM of compliant parts

90

CHAPTER 6

Covariance Calculation and


Statistical FEA Solution

6.1 Introduction
As stated in Chapter 1, the objective of the FASTA process is to statistically
predict the force, deformation, and stress produced within compliant parts by assembly
processes. Statistical representations of a system include both a mean and a variance.
However, only the mean solution for the stiffness equation has been treated thus far
(equation 4-1). The statistical solution requires the determination of the covariance of the
system. The covariance of a linear system may be obtained directly by the following
matrix operation. Given a linear matrix equation:

{Y} = [M ]{X}

(6-1)

The covariance of {Y}, written [Y ] , is related to the covariance of {X}, [ X ] , by the


following equation [Johnson 1988]:

[ Y ] = [M ][ X ][M ]T

(6-2)

This statistical property allows for the statistical representation of the relation between
force and displacement. If {F} = [K ]{ } describes the relationship between force and
deflection of a structure, then the statistical description is expressed by:
91

Mean : { F } = [K ]{ }

Covariance : [ F ] = [K ][ ][K ]

(6-3)

where { F } and { } are the mean force and deflection vectors, respectively, and [ F ]
and [ ] are the corresponding statistical covariances.
In equation 6-3, [ ] denotes the gap covariance matrix. The steps required to
calculate this covariant matrix are presented in this chapter. Using equation 6-3, the
statistical solution for the closure force along the gap can then be determined, and used to
find internal stresses and deformations in a compliant assembly. These calculations
describe Steps 3 and 4 of the FASTA process: the covariance calculation and the
statistical FEA solution. In this chapter these steps are discussed in detail and illustrated
using the example assembly introduced in Chapter 4.

6.2 Step 3: Covariance Calculation


In Step 3 of FASTA, the gap variation components calculated in Step 1 (Chapter
4) are used to solve for the gap covariance matrix [ ] . An overview of Step 3 is shown

{ }

{ }

in Figure 6-1. Inputs Ti2 and Ri2 are the variance vectors of the rigid body
translation and rotation, which occur in the gap due to the accumulation of process

{ }

2
variations in assembly. Vector SPi
is the gap variation due to local distortion of the

mating surfaces. The variance vectors are obtained by tolerance stackup analysis or from
measurement of production parts. A surface continuity model is also needed in this step
to predict the interdependence between points on the mating surfaces.

92

FASTA Step 3 Overview

{ } { }
2
Ti

2
Ri

Surface
Continuity
Model

{ } Inputs
2
SPi

3. COV

[ ] Output
Figure 6-1 Step 3 process summary

In this section covariance is introduced, surface continuity models are presented, and a
hybrid method for obtaining the surface covariance by combining different surface
models is discussed. Procedures for introducing rigid contributions to the gap covariance
matrix are also presented. As a demonstration, the gap covariance for the example
assembly is determined and the process is summarized.

6.2.1 Introduction to covariance


Covariance can be defined simply as the interdependence or correlation between
variables. Figure 6-2 illustrates this idea using points located by variables x and y. As
shown in this figure, uncorrelated variables are independent, partially correlated variables
show some dependence, and fully correlated variables are completely dependant.

93

Uncorrelated

Partially correlated

Fully correlated

Figure 6-2 Example of correlation between variables x and y

With varying degrees of interdependence, covariance provides a tool for predicting the
behavior of one variable given a change in another.
When modeling flexible parts, there are two types of covariance that must be
included: material covariance and geometric covariance [Merkley 1998]. Material
covariance results from the internal elastic material interactions within a compliant part,
and is accounted for by applying the stiffness matrix [K] appropriately. Geometric
covariance [ GP ] accounts for surface continuity at the mating contact points of a part P.
This is where the interdependence between fastening points is modeled (assumed
independent in Step 1 of FASTA). Geometric covariance can either be calculated
directly from a set of measured part data, or predicted using surface continuity models
(discussed in the next section). For mating surfaces A and B the gap covariance
matrix, [ ] , is calculated using the following equation:

[ ] = [ GA ] + [ GB ] + [ Trans ] + [ Rot ]

(6-4)

In this equation matrices [ GA ] and [ GB ] are the geometric covariances for the mating
surfaces of parts A and B, and matrices [Trans ] and [ Rot ] represent the rigid covariance
94

contributions (translation and rotation) to the gap covariance. Each of these matrices are
square with n rows and n columns, where n is equal to the number of degrees of freedom
at the fastening points along the gap.
While each of the matrices in equation 6-4 may have off-diagonal terms,
indicating interdependence, summing them linearly assumes that the individual
covariance matrices are independent of each other. The calculation of geometric
covariance matrices and rigid covariance matrices are treated in detail in the next two
sections.

6.2.2 Surface continuity models


In the product development cycle, analysis and design improvements are much
more cost effective early in the design, rather than during production. Since statistical
part data is not available in the design phase, predictive methods are required for
modeling surface characteristics of flexible parts in order to estimate the geometric
covariance. Even in predictive stages, some surface variation information must be known
or inferred in order to create the necessary models. Production processes have unique
characteristics and the resulting surface variation will depend on the process, tooling and
material. This information could be contained in a database of surface models for
specific manufacturing processes. In this thesis, such information is assumed to be
available.
Using sinusoids and polynomial series, [Tonks 2002] presented a robust method
for modeling surface continuity for use in FASTA. The sinusoidal and polynomial series
models for calculating geometric covariance are presented in this section, as well as a
95

hybrid model that allows for the combination of multiple models to meet certain
conditions. The example assembly is used to illustrate the analysis. For derivations and
detailed treatments of these methods, refer to [Tonks 2002].
The objective of these methods is to describe surface variation in terms of random
surfaces, rather than sets of random points. The covariance matrix for a set of
independent random points is diagonal (zero covariance). The sinusoidal and polynomial
methods, described in the following sections, are the means of deriving the necessary
continuity conditions for obtaining the off-diagonal covariance terms.

6.2.2.1 Sinusoidal model


Given a vector {Y } containing a set of N independent points with a standard

{}

deviation of y , it is possible to create a vector Y whose points have the same standard
deviation, but are constrained to lie along a specified sinusoid. An example of this curve
fit is shown in Figure 6-3.

Figure 6-3 Sinusoidal curve fit of independent points [Tonks 2002]

96

{}

The relationship between Y and {Y } can be written using a transformation matrix [S S ]


with the following equation:

{Y} = [S

]{Y }

(6-5)

The subscript S in the matrix denotes that the result follows a prescribed sinusoid. This
transformation matrix provides a link between independent points and points along a
continuous curve. Using this relation, the geometric covariance for a part surface that can
be modeled using sinusoids is:

[ GP ] = [S S ][ SP ][S S ]T

(6-6)

Where [ SP ] is the surface variation constraint for the independent points along the
mating surface of part P. This diagonal matrix is created by simply multiplying the
surface variance vector for part P (obtained in Step 1 of FASTA) by the identity matrix,
as shown in the following equation:
2
}[I ]
[ SP ] = { SPi

(6-7)

A general surface may be described by a series of sinusoids of differing frequency


and amplitude. For a series of sinusoids, described by a set of M frequencies { f } with
corresponding amplitudes, or weighting coefficients {a}, the transformation matrix can be
solved for the ith row and the jth column as:
S S ,ij =

M 1

k =0

2f k
2
( j i)
a k ,norm cos
N
N

97

(6-8)

where the normalized weighting coefficients ak ,norm are calculated from the given set of
weighting coefficients by the following equation:
a k ,norm =

k
M 1

a
k =0

(6-9)
k

It should be noted that the order of the transformation matrix is always the same as the
order of the matrix that it is multiplying. In this analysis, this will always be n rows and
n columns, corresponding to the degrees of freedom of the gap assembly points. With
this solution for the transformation matrix, the geometric covariance for surfaces modeled
as sinusoids can be calculated using equation 6-6.

6.2.2.2 Polynomial series models


Another versatile way to model surface continuity is with polynomial series.
Using polynomial series that are complete orthogonal polynomials, such as Chebychev
and Legendre polynomials, allow for Fourier series simplifications and are ideal for
surface modeling. Equation 6-5 can be modified to use a polynomial series
transformation matrix, resulting in a set of points that maintain the random standard
deviation, but are constrained to follow a specified polynomial series. The geometric
covariance transformations for Chebychev and Legendre polynomials are denoted as
[ S ch ] and [ S lg ] respectively. Plots of these two orthogonal series are shown in Figure
6-4 for the 0-4th orders.

98

Chebychev Polynomials

Legendre Polynomials

Figure 6-4 Orthogonal polynomial plots [Tonks 2002]

Chebychev and Legendre polynomials are presented and applied to geometric covariance
in this section.

Chebychev Polynomials:
Chebychev polynomials are a set of polynomials of increasing order, which have
been normalized to have a range of 1 xi 1 , and are bounded in the y-direction by 1
and -1. Chebychev polynomial orders zero through three are written as:
T0 ( xi ) = 1
T1 ( xi ) = xi

(6-10)

T2 ( xi ) = 2 xi2 1
T3 ( xi ) = 4 xi3 3xi

Any set of x-values, {u}, with N terms can be normalized to fit within the limited range of
Chebychev polynomials using the following equation:

xi =

ui

1
(N + 1)
2

(6-11)

1
(N 1)
2

99

Given a series of polynomials with coefficients {a} for each polynomial order up to M,
the Chebychev geometric covariance transformation matrix can be solved for the ith row
and the jth column as:
S ch,ij =

1
a 0,normT0 (x j ) +
N

2
N

M 1

a
k =1

T (x j )Tk ( xi )

k , norm k

(6-12)

Again the weighting coefficients are normalized using equation 6-9. With the sensitivity
matrix solution in equation 6-12, the geometric covariance for surfaces modeled using
Chebychev polynomials can be calculated as:

[ GP ] = [S ch ][ SP ][S ch ]T

(6-13)

Legendre Polynomials:
Legendre polynomials are also defined over the range of 1 xi 1 and bounded
at the ends in the y-direction by 1 and -1. Unlike Chebychev, Legendre polynomials only
reach values of 0.5 at their inflection points instead of 1 (see Figure 6-4). Orders
zero through four of Legendre polynomials are written as follows:
P0 ( xi ) = 1
P1 ( xi ) = xi

1
3 xi2 1
2
1
P3 ( xi ) = 5 xi3 3xi
2
1
P4 ( xi ) = 35 xi4 30 xi2 + 3
8
P2 ( xi ) =

(6-14)

Again equation 6-11 is needed to normalize any set of points that fall outside of the
defined range for this polynomial set. A given a set of weighting coefficient {a} for
100

each polynomial order up to M, must be normalized using equation 6-9. The Legendre
geometric covariance transformation matrix can be calculated at the ith row and jth
column by:
S lg,ij =

M 1

l =0

2l + 1
a k ,norm Pl (x j )Pl ( xi )
N

(6-15)

After solving for the sensitivity matrix in equation 6-15, the geometric covariance for
surfaces modeled using Legendre polynomials can be calculated as:

[ GP ] = [S lg ][ SP ][S lg ]T

(6-16)

6.2.2.3 Hybrid method


The objective of the covariance methods is to include the effects of surface
waviness wavelength in the predicted assembly force, stress and displacement variations.
[Soman 1999] observed three ranges of variation, each with distinctly different effects,
which he described in terms of the wavelength-to-part length ration /L. Table 1-1
describes the three ranges.
Table 6-1 Ranges of surface variation and their effects

Wavelength Description
Effect
Large scale warping Dominant, when present
/L 1
Moderate
1/6 /L < 1 Surface waviness
Surface roughness
Negligible
/L < 1/6
This diminishing effect as wavelength gets shorter and shorter is partly due to the fact
that the amplitude generally diminishes for short wavelengths
[Soman 1999] determined that for geometric covariance modeling, the spectral
analysis method presented by [Bihlmaier 1999] were limited to wavelength-to-part length
101

ratios of 1 or less (/L 1). Based on a similar concept, the sinusoidal method presented
by [Tonks 2002] is also subject to this limitation. The polynomial series methods,
however, best model covariance for wavelength-to-part length ratios of 1 or more (/L
1). Table 6-2 summarizes the useful range for the surface continuity models presented in
the last section based upon /L.
Table 6-2 Surface continuity model use range

Surface Continuity Model Use range


Sinusoidal
/L < 1
Polynomial Series
/L 1
closer
to 1
Chebychev
/L >>1
Legendre
In order to take advantage of the overlapping surface continuity ranges shown in
this table, [Tonks 2002] created a hybrid method to combine multiple continuity models
into a single geometric covariance. In this way, a surface that contains both short and
long wavelengths (relative to surface length) can be modeled using a combination of
sinusoidal and polynomial series models. Given weighting coefficients a and b, the
hybrid geometric covariance for a part P,[ GPH ] , can be written as:

[ ]
GPH

b2
a2
GPS
GPPoly + 2
= 2
2
2
a +b
a +b

(6-17)

In this equation, [ GPPoly ] is the polynomial geometric covariance for frequencies where
/L 1 (using Chebychev or Legendre series) and [ GPS ] is the sinusoidal covariance
for the frequencies with /L < 1. Equation 6-17 can be adapted for any number and type
of covariance models, allowing for modeling versatility. To combine more than two

102

geometric covariance models, the normalized weighting coefficients are calculated using
the following equation.

c norm =

ci2
k

c
i =0

(6-18)
2
i

6.2.3 Rigid contribution to gap covariance


As shown in equation 6-4, the gap geometric covariance includes not only the
mating surface geometric covariances, but also the rigid covariance contributions from
assembly. These rigid contributions can be separated into translational and rotational
components, which describe the covariant rigid body modes of the overall gap
covariance. In this section, the translational and rotational covariance matrices are
defined and calculated using rigid variance outputs from Step 1 of FASTA and the
transformation matrices derived in the previous sections.

Translational Covariance:
Rigid translational variation seen at the gap is a result of the statistical
accumulation of part variation in the assembly. Using the DLM method in Step 1 of
FASTA, the magnitude of the translational variation at the gap is calculated by including
only variations that contribute to translation. It should be noted that this rigid form of
variation is completely covariant, meaning that the movement of one point along the
surface results in an equal displacement of the entire surface. This is illustrated in Figure
6-5 as the vertical movement of a horizontal line.

103

Upper limit
Mean

Variation
Lower limit

Figure 6-5 Rigid translational variation

{ }

Given the translational variance at each node i along the gap, Ti2 from step 1 of
FASTA, the covariance of the gap translation can be found as:

[ Trans ] = [S ][ T ][S ]T

(6-19)

{ }

In this equation, [ T ] is a diagonal matrix created using Ti2 along the diagonal, and

[S ] is the transformation matrix that defines the covariance of the translation.

By looking

at the polynomial series models in figure 6-4, it is apparent that a zero-order Chebychev
or Legendre polynomial can be scaled to model any horizontal line. Using the Legendre
series, the transformation matrix for translational variation can be created using equation
6-15, where the weighting coefficient vector is simply:

{a} = {1 , 0}

(6-20)

This means that only a zero-order term is present and it is scaled by 1. Equation 6-19
then becomes:

[ Trans ] = [S (0)lg ][ T ][S (0)lg ]T

(6-21)

Where [ S (0 )lg ] denotes a zero-order Legendre transformation. It should be noted that if


the translational variance is the same at each node, the resulting translational covariance

( )

matrix is completely covariant with the same value Ti2 throughout the matrix. This
104

matches the model in Figure 6-5, where the movement of one point along the surface
results in an equal displacement of the entire surface.

Rotational Covariance:
Rigid rotational variation seen at the gap is also a result of the statistical
accumulation of part variation in assembly. Again from the DLM method in step 1 of
FASTA, the rotational variation at the gap is calculated by including only part and
assembly variations that contribute to rotation along the gap. The resulting gap variation
is limited to max by the characteristic length of the gap and the maximum vertical
difference at the surface ends. This variation is illustrated in Figure 6-6 by a linear line
that rotates about the center. Recall from equations 4-17 and 4-18, that the rotational
variation can either be represented by a constant angular value, or converted to
translational variation that increases linearly from the center of the surface.
Variation
max

Mean

Figure 6-6 Rigid rotational variation

{ }

With the rotational variance at each node i along the gap, Ri2 from step 1 of
FASTA, the covariance of the gap rotation is determined by:

[ Rot ] = [S ][ R ][S ]T

(6-22)

{ }

Where: [ R ] is a diagonal matrix created using Ri2 along the diagonal, and [S ] is the
transformation matrix that defines the covariance of the rotation. Again looking at the
polynomial series models in Figure 6-4, it is apparent that a first-order Chebychev or
Legendre polynomial can be scaled to model any linear line. With the Legendre series,
105

the transformation matrix for rotational variation can be calculated using equation 6-15
where the weighting coefficient vector is now:

{a} = {0 , 1}

(6-23)

This means that only a first-order term is present, which is scaled by 1. With this scaling,
the slope of the first-order polynomial is also 1, and the evaluation of equation 6-22
results in a final scaling of the linear covariance model by the rotational variance
contained along the diagonal of matrix [ R ] . This is allowable since the angular
magnitude of the rotation is sufficiently small that the slope of the line representing
rotational variance is simply the rotational variance (small angle theorem).
Using a first order polynomial to create the transfer function, equation 6-22 can
now be written as:

[ Rot ] = [S (1)lg ][ R ][S (1)lg ]T

(6-24)

Where [ S (1)lg ] represents a first-order Legendre transformation. The rotational


covariance matrix is saddle-shaped, with maximum values in the upper-left and lowerright corners. Corresponding negative values are found in the upper-right and lower left
corners. An example of this is given in the next section.

6.2.4 Gap covariance of the example assembly


With solution methods presented for calculating the geometric and rigid
covariance components, the gap covariance matrix can be solved using equation 6-4. In
this section the solution process is illustrated using the example assembly from Chapter 4.
In addition to the gap variance vectors calculated in Step 1 of FASTA, a surface
106

continuity model for each gap surface is also required as an input. For the example
assembly, the surface continuity models used in the analysis are presented in Table 6-3.
Table 6-3 Surface continuity models for example assembly

Part A
Frequency Components (/L):
Weighting Coefficients:
Part B
Frequency Components (/L):
Weighting Coefficients:

[0.1, 1, 7]
[8, 2, 1]
[0.2, 2, 8]
[5, 2, 2]

In the covariance calculation, the first task is to solve for the geometric covariance
for both of the mating surfaces of the gap, then combine them with rigid covariance
contributions to form the gap covariance matrix. Comparing the frequency components
of parts A and B in Table 6-3 with Table 6-2, it is apparent that the hybrid method is
needed to model both the low and high wavelength-to-part length ratios. In this way the
Legendre model can be used for components with /L greater than 1, and the sinusoidal
model can be used for all other components. The weighting coefficients given in Table
6-3 dictate how these are combined. Using the surface variances in equation 4-36 and the
hybrid method, the geometric covariance matrix solutions for parts A and B are shown
graphically in Figure 6-7. While the shapes of these matrices are similar, geometric
covariance of the gap surface on part B is nearly an order of magnitude smaller than that
of part A.

107

a) Geometric covariance of part A

b) Geometric covariance of part B

Figure 6-7 Geometric covariance solutions for example assembly gap surfaces

The next task is to solve for the gaps rigid covariance components. The
translational covariance for the example assembly is calculated using equation 6-21 and
the translational variance vector solution in equation 4-18. The gap rotational covariance
is determined using the rotational variance vector solution in equation 4-23 with equation
6-24. The rigid covariance solutions for the example assembly are shown in Figure 6-8.

a) Gap translational covariance

b) Gap rotational covariance

Figure 6-8 Rigid gap covariance solutions for example assembly

108

From this figure it is apparent that translation is a much larger contributor to the overall
gap covariance than the rotation. As discussed earlier, the shape of the rigid covariance
matrices in Figure 6-8 match their respective modeling assumptions with the completely
and equally covariant translation being much different in appearance than the saddleshaped covariance from rotation.
The final task in the calculation of the gap covariance matrix is to sum the
different covariance components as shown below (from equation 6-4):

[ ] = [ GA ] + [ GB ] + [ Trans ] + [ Rot ]

(6-25)

For the example assembly the final gap covariance solution is displayed graphically in
Figure 6-9a. A plot of the individual covariance diagonals (for each component in
equation 6-25) is shown in Figure 6-9b for comparison. Looking at this plot, it is
apparent that the translational covariance has the largest contribution to the final shape,
while the rotational has the smallest.

a) Gap covariance solution

b) Diagonal of covariance components

Figure 6-9 Gap covariance matrix solution and contributing components, for example assembly

109

6.2.5 Step 3 process summary


In this section, Step 3 of the FASTA process, the covariance calculation, has been
treated in depth. Geometric covariance was introduced and accounted for using surface
continuity models based upon sinusoids, polynomial series, or a hybrid combination of
multiple models. Rigid covariance components, translation and rotation, were also
determined using variance inputs and polynomial models. The output of this step, the
gap covariance, was solved for by combining the various covariance contributions.
Finally the process was illustrated using the example assembly from Chapter 4. A
summary of Step 3 of the FASTA process is listed in Table 6-4 and shown illustrated
with a process flow diagram in Figure 6-10.

110

Table 6-4 Summary of steps to calculate gap covariance

{ }{ }{ }

2
Inputs: Ti2 , Ri2 , SPi
Outputs: [ ]
Analysis Steps:

I. Covariance Component Solutions


A. Geometric covariance of flexible surfaces
1. Create transformation matrix using surface continuity model
2. Create diagonal matrix using surface variation variance vector
3. Calculate geometric covariance for surface of part P
4. Repeat for additional flexible part surfaces at gap
B. Translational Covariance (Rigid)
1. Create transformation matrix using zero-order Legendre polynomial
2. Create diagonal matrix using translational variance vector
3. Calculate translational covariance of gap
C. Rotational Covariance (Rigid)
1. Create transformation matrix using first-order Legendre polynomial
2. Create diagonal matrix using rotational variance vector
3. Calculate rotational covariance of gap
II. Gap Covariance Calculation
1. Sum geometric, translational, and rotational covariance components
2. Output solution

111

FASTA Step 3: Covariance Calculation


Surface Cont. Model

Sinusoidal
Polynomial Series
Legendre
Chebychev
Hybrid (mixed-model)

Inputs

{ }
2
SPi

Transformation
Matrix: [S ]

Geometric
Covariance

[ SP ]

Calculate
Geom. cov.

[ GP ]

Repeat for each compliant gap surface

Translational
Covariance

{ }
2
Ti

[ T ]

Transformation
Matrix: [ S (0 )lg ]

Calculate
Transl. cov.

[ Trans ]

[ ]
Output

Rotational
Covariance

{ }
2
Ri

[ R ]

Transformation
Matrix: [ S (1)lg ]

Calculate
Rotatl. cov.

[ Rot ]

Figure 6-10 Process flow diagram for calculating gap covariance

112

6.3 Step 4: Statistical FEA Solution


The final step of the FASTA method, step 4, uses the gap covariance matrix [ ] ,
the mean gap vector { }, the equivalent stiffness matrix[K eq ] , and FEA to statistically
solve for: assembly force, internal deformations and stress. An overview of this process
is shown in Figure 6-11.

FASTA Step 4 Overview


[ ]

{ }

[K eq ]

Inputs

4. Statistical
FEA
FEA
F( , )
( , )

Outputs

Stress( , )

Figure 6-11 Step 4 process summary

In this section the assembly force is statistically solved, the mean of the deformation and
stress is calculated, the covariant equations for deformation and stress are presented, and
the process is summarized. Each step is illustrated using the example assembly.
113

6.3.1 Statistical solution of assembly force


A statistical prediction of assembly force for compliant parts is useful both when
designing assembly tooling and examining the load conditions on the parts themselves.
Replacing the stiffness in equation 6-3 with the equivalent stiffness between mating parts,
the mean and covariance of the assembly force are:
Mean : { F } = [K eq ]{ }

[ ] [ ]

Covariance : [ F ] = K eq [ ] K eq

(6-26)

Steps 1-3 of FASTA provide the inputs to these equations, requiring only simple matrix
algebra to evaluate.
With the example assembly calculations presented in previous sections and
chapters, the predicted mean and variation of the assembly force at the gap can be
calculated due to both rigid and compliant variation sources. Since the mean gap at each
mating point was zero (no mean separation), the mean closure force is also zero at each
point. For illustrative purposes, suppose that the mean gap at every fastening location
was actually 0.25 inches, perhaps the result of a defect in the tooling. The mean gap
vector would then be:

{ } = {0.25

0.25 0.25 0.25 0.25}

(6-27)

Using equation 6-27 and the equivalent stiffness for the example assembly, the mean
solution for the assembly force from equation 6-26 would be:

{ F } = {0.032

0.021 0.030 0.021 0.032} lbf


T

114

(6-28)

Independent of the mean value, the predicted force covariance is calculated by


solving the second part of equation 6-26. Using the equivalent stiffness and gap
covariance determined for the example assembly, the resulting force covariance solution
is shown graphically in Figure 6-12.

Figure 6-12 Force covariance matrix for example assembly

Taking the square root of the force variance at each node (found along the diagonal of the
covariance matrix) a vector of standard deviations of the force at each fastening point can
be determined:

{ F } = {0.981

2.378 2.921 2.378 0.981} lbf


T

(6-29)

Multiplying equation 6-29 by three gives an estimate of the 3 range of closure force
seen at each node by 99.7% of assembled parts at the gap. In other words, 99.7% of the
assemblies will require 8.76 pounds of force or less to close the misaligned gap.
Comparing equations 6-28 and 6-29, it is apparent that the variance of the closure force is
significantly higher than the mean closure force. Much of the difference is due to the
115

surface waviness of the mating parts accounted for in the gap covariance, which
[Bihlmaier 1999] proved to significantly increase closure forces and stress.

6.3.2 Mean solution of internal deformation and stress


In order to determine the mean solutions for deformation and stress in each
flexible part, the displacement seen at each fastening node must first be calculated. It
should be noted that each compliant part is to be treated separately for this step. From
equations 5-11 and 5-12 the mating edge boundary displacements for part A {A } (mated
to part B) can be calculated from the mean gap vector using the following equation:

{ A } = [K RA ]{ }

(6-30)

Where the stiffness ratio for part A is defined as:

[K RA ] = ([K red A ] + [K red B ])1 [K red B ]

(6-31)

In this equation [K red A ] and [K red A ] are simply the reduced stiffness matrices for parts A
and B. The boundary displacements for part B { B } are calculated the same way, but
switching the Bs and As in equations 6-30 and 6-31. For example parts A and B, the
boundary displacement solutions at equilibrium (assuming the mean gap displacement in
equation 6-27) are:

{ A } = {0.2478
{ B } = {0.0022

0.2515 0.2485 0.2515 0.2478}

0.0015 0.0015 0.0015 0.0022}

(6-32)

In this displacement solution, the effects of the relative stiffness between parts A and B
can be seen, with most of the equilibrium displacement taking place in more flexible part
116

A. Positive values in equation 6-32 correspond to displacement in the direction of


closure for the gap being analyzed.

6.3.2.1 Mean deformation solution


With the mean displacements along the mating boundary known, the next step is
to solve for the mean displacements seen throughout each parts interior nodes. The
complete mean displacement solution for each part can be obtained from the mean
stiffness equation containing every degree of freedom:

{ F }c = [K ]c { }c

(6-33)

In this equation, the subscript c is used to denote a completely populated matrix


containing all degrees of freedom in the part under consideration. The mean
displacement vector cannot be solved directly using equation 6-33 since the complete part
stiffness matrix [K ]c is singular and cannot be inverted. Instead, equation 6-33 must be
partitioned (as shown in equation 5-5) and solved in parts using the boundary conditions.
This process is shown in Chapter 5, section 5.3.1. From equation 5-8, the equation for
the mean interior displacements can be written in terms of the mean boundary
displacements as:

{ i } = [K ii ]1 [K ib ]{ b }

(6-34)

Again [K ii ] is the interior node stiffness and [K ib ] is the stiffness interaction between the
interior and boundary nodes. The complete mean displacement vector can be
reconstructed from the mean interior and boundary displacements:

117

{ }c


= b
i

(6-35)

The challenge with the solution presented above is one of keeping track of node
locations in partitioning and reconstruction. Since a fully developed FEA model has
already been created for each part in step 2 of FASTA, it is much simpler to apply the
known mean boundary displacements in commercial FEA software and solve for the
mean part deformation. An example of this is shown in the next section.

6.3.2.2 Mean stress solution


With mean nodal displacements calculated, the mean stresses can be evaluated
one element at a time using the finite element relation [Bihlmaier 1999]:

{ e } = [D e ][B e ]{ e }

(6-36)

In this equation {e } is the element mean stress vector and {e } is the nodal mean
displacement vector. The material elastic properties are contained in [De ] , called the
element constitutive matrix, and the partial derivatives of the element shape functions are
contained in [B e ], called the element kinematic matrix.
The solution to equation 6-36 is contained within any commercial FEA software
package and can be solved along with the interior nodal displacements, given the
boundary conditions. With the mean boundary displacements in equation 6-32 applied in
ANSYS, combined plots of mean deformation and stress are shown in Figure 6-13 for
example parts A and B.

118

a) Mean deformation and stress: Part A

b) Mean deformation and stress: Part B


Figure 6-13 Plots of mean deformation and stress for example parts in ANSYS

119

6.3.3 Covariance equations for deformation and stress


Covariance calculations for deformation and stress involve significantly more
computation than the mean calculations presented in the last section. This is not
surprising, however, since the resulting covariance solutions describe the
interdependence between every node in a part. As such, these solutions provide a
statistical view of how tolerance stack-up affects the deformation and stresses seen by
flexible parts in assembly. It is for this reason that the fourth step of FASTA is referred
to as statistical FEA. In this section the processes for determining the covariant
deformation and stress are introduced. For a more detailed treatment of the topic see
[Bihlmaier 1999].

6.3.3.1 Covariant deformation solution


In order to solve for the standard deviation of assembly deformation in a
compliant part using FEA, a covariant solution is required to account for the
interdependence between each node. Returning to the relationship between force and
displacement in equation 6-3, the covariance equation can be inverted to solve for the full
displacement covariance matrix:

[ ]c = [K ]c1 [ F ]c [K ]c1

(6-37)

Again the subscript c denotes completely populated or global matrices relating to all
degrees of freedom in a part. With applied forces only present at mating boundary nodes,
the complete force covariance matrix [ F ]c is zero for all terms other than those

120

corresponding to the gap force covariance matrix solved for using equation 6-25. These
can be easily mapped into corresponding node locations within the matrix.
As stated before, the complete part stiffness matrix [K ]c is singular and cannot be
inverted without first applying known displacement constraints. Since no displacement is
possible at completely constrained nodes, the covariance with relation to these nodes is
simply zero. This means that the part displacement covariance matrix [ ]c contains
rows and columns of zeros corresponding to constrained nodes. By removing these
constrained degrees of freedom from each of the matrices in equation 6-36, the stiffness
matrix is made nonsingular and equation 6-36 can be solved.
Once the part displacement covariance matrix [ ]c is calculated, the variances
corresponding to each node in the FEA model can be pulled from its diagonal. By
mapping the nodal displacement variances back to corresponding part coordinates, a
displacement variance surface plot can be made. This plot provides a clear picture of
how assembly variation affects the statistical displacement of a flexible surface.
To illustrate the covariant deformation solution using a displacement variance
surface plot, equation 6-36 was applied to example part A and solved. Nodal variances in
the y-direction were then mapped to corresponding x-z locations of the FEA model
(Figure 6-14a) for the cantilevered (horizontal) surface of part A. The resulting
displacement variance surface plot is shown in Figure 6-14b. Due to rigidity in the ydirection, the vertical surface of part A deflects very little and is omitted from the plot.

121

a) FEA model of part A

b) Displacement variance surface plot

Figure 6-14 Displacement variance surface plot of the horizontal surface of example part A

As anticipated, the plot in Figure 6-14b suggests that the variance of the displacement
along the flexible surface increases toward the mating end of the part. The quantification
of the surface variance allows for a statistical prediction of the behavior of a population
of assemblies, and is a major analysis contribution of the FASTA process.

6.3.3.2 Covariant stress solution


Unlike the mean stress solution, the covariant stress solution cannot be solved one
element at a time. Since covariance exists between all nodal displacements in a part, it
also exists between all nodal stresses. Treating one element at a time does not account
for this interdependence of nodal stress variation, requiring a more complex solution. In
order to obtain the variance of stresses seen in a part, a stress covariance equation
including all degrees of freedom has to be solved. This equation was presented by
[Bihlmaier 1999] as:

[ ] = [D][B][ ][B]T [D]T


122

(6-38)

where: [ ] is the global stress covariance matrix, [ ] is the global displacement


covariance matrix, [D] is the global constitutive matrix, and [B ] is the global kinematic
matrix. Detail into the solution of equation 6-37 was given by [Bihlmaier 1999], and will
not be treated in this work.
While the method for the calculating the stress covariance matrix [ ] exists, its
implementation on complex geometry requires extensive programming that is beyond the
scope of this thesis. It is suggested that such an automated implementation could be
created as a third-party interface to an FEA program like ANSYS, allowing for the full
application of statistical FEA. This is a step where future work must be done to use
FASTA to its full potential.

6.3.4 Step 4 process summary


The fourth and final step of the FASTA process, the statistical FEA solution, was
presented in this section and illustrated using the example assembly from Chapter 4.
Assembly forces at the gap fastening points were solved statistically, and the mean
deformation and stress were calculated. Covariant solutions for deformation and stress
were also presented. The displacement variance plot was introduced as a means of
representing displacement variation in assembled compliant parts. Step 4 of the FASTA
process is summarized in Table 6-5 and illustrated with a process flow diagram in Figure
6-15.

123

Table 6-5 Summary of steps for statistical FEA solution

Inputs: [ ] , { }, [ K eq ]
Outputs: F( , ), ( , ), Stress( , )
Analysis Steps:
A. Statistical Solution of Assembly Force
1. Calculate mean closure force
2. Calculate force covariance matrix
B. Mean Solution for Part Deformation and Stress (individual part)
1. Determine equilibrium displacement of part for mean gap
2. Apply mean part displacement as load condition in FEA
3. Solve FEA for load condition and output mean displacement and stress
4. Repeat for each compliant part at gap
C. Covariant Solutions for Part Deformation & Stress (individual part)
1. Obtain full part stiffness matrix from FEA
2. Create global force covariance matrix using boundary force covariance
3. Remove fixed degrees of freedom from global matrices
4. Solve for part displacement covariance matrix
5. Determine global constitutive and kinematic matrices from FEA
6. Solve for part stress covariance matrix
7. Repeat for each compliant part at gap

124

FASTA Step 4: Statistical FEA Solution


Inputs
{ }

Outputs
Calculate
mean force

F( )

Assembly
Force Solution

[K eq ]

[ ]

{ F }

Calculate
force cov.

[K R ]

[ F ]

F()

Mean
Solutions

Determine
mean part
displacement

()
FEA
Stress( )

Repeat for flexible parts

FEA

[K ]c

Covariance
Solutions

Create
global force
cov. matrix

[ F ]c
Remove
constraint
dof

[ ]c

[D]
FEA

[B]

Solve for
stress cov.

()

[ ]

Repeat for flexible parts


Figure 6-15 Process flow diagram for statistical FEA soltion

125

Stress()

6.4 Chapter Summary


In this chapter, statistical covariance was used to present the lat two steps of the
FASTA process: the covariance calculation, and the statistical FEA solution. Each of
these steps was discussed in detail and summarized in tabular and process flow format.
This concludes the defining of the individual steps of the FASTA process.

126

CHAPTER 7

Form Variation in
Flexible Assemblies

7.1 Introduction
Up to this point in the thesis, form variations at flexible joints other than the gap
have been ignored. The purpose of this chapter is to look at how form variations at these
joints can propagate through an assembly and affect gap variation. Form variation
modeling at rigid joints is first reviewed and then compared with flexible joint variation
propagation. A method for using FASTA at each flexible joint to account for form
variation is proposed, and modes of variation propagation are discussed.

7.2 Review of Rigid Form Variation Modeling


In considering the propagation of form variation through mating joints in flexible
assemblies, it is useful to first look at rigid joints as a modeling foundation. As discussed
in Chapter 3, form variation propagates in an assembly only at the surface contacts or
joints between parts. For rigid assemblies, form variation is modeled in the DLM using
zero-length vectors applied at the point of mating contact. Although zero in length, each
vector still has direction and variation. Based upon geometric feature tolerance and the
joint type along the axis under consideration, form variation can propagate either as
127

translation or rotation (see Figure 3-3). Such rigid variance propagation accumulates
statistically through assembly.
An example of how form variation propagates rigidly can be seen in Figure 7-1,
where geometric feature tolerances are applied to rigid groove and cylinder surfaces.
Translation

A
B
y

Kinematic
adjustiment

x
Translation

a) Geometric feature tolerances

b) Joint propagation of variation

Figure 7-1 Example of rigid form variation propagation

Each of the two joints in this assembly is the result of the cylinder contacting a planar
surface inside of the groove. As discussed in chapter three, this is called a cylindrical
slider joint. Referring to Table 3-2, it can be seen that both the flatness tolerances and the
circularity tolerance result in translational variation normal to the surface contacts. This
is illustrated by two variation vectors at each of the joints in Figure 7-1b, representative
of translational variation components from both parts. These translational variations
result in a kinematic adjustment of the assembly dimensions, changing the distance
between the center of the cylinder and the bottom of the groove.
Looking only at the 2-D case for the assembly in Figure 7-1, the form variation
components of the kinematic constraint equation can be written as:

128

H x (form) = 1 cos(90 + ) + 2 cos(90 + ) + 3 cos(90 ) + 4 cos(90 )


H y (form) = 1 sin (90 + ) + 2 sin (90 + ) + 3 sin (90 ) + 4 sin (90 )

(7-1)

where the s are summarized in Table 7-1. The form variation contribution is then
combined with the dimensional variation to determine the kinematic or assembly
variations as described in chapter 3. From this example it is apparent that form variation
in rigid assemblies is inserted at the joints and propagates rigidly, based upon joint type.
Table 7-1 Geometric tolerance variable for groove and cylinder example

Variable

Description

Nominal

Variation

Variation from the flatness on


the angled surface
Variation from the circularity
at the angled joint
Variation from the circularity
at the lower joint
Variation from the flatness on
the horizontal surface

A/2

B/2

B/2

C/2

2
3
4

7.3 Flexible Parallel to Rigid Form Variation Model


Just as for rigid assemblies, form variation in flexible assemblies also acts at
assembly joints. There is, however, a fundamental difference between rigid joints (with
both mating parts rigid) and flexible joints (with at least one flexible part). Due to their
compliant nature, flexible parts can deform and thus accommodate for mating surface
variation. This is illustrated by a comparison of rigid and flexible joints in Figure 7-2.

129

Flexible
Sheet

1
Rigid Block

a) Example of rigid joint

Fasteners

b) Example of flexible joint

Figure 7-2 Comparison of rigid and flexible joints

In this figure, the flexible sheet behaves similarly to the rigid block before fastening.
Once fastened, however, the initial rotation of the sheet is absorbed partially by distortion
as the part conforms to the rigid mating surface.
With deformation of compliant parts at flexible assembly joints come several
results. From Figure 7-2 b, it is apparent that the rigid component of rotation at the joint
has been altered as a result of part deformation. Thus, rigid form variation propagation is
modified by part compliance. Distortion, however, not only stores energy, but transfers it
in the form of residual stresses and elastic deformation. These, too propagate through
flexible parts and can affect the assembly. With this discussion, two challenges in
modeling form variation propagation in flexible assembly have been brought to light.
These are:
1. Quantifying the variation due to part distortion,
2. Predicting the effects of elastic deformation on the assembly.
The purpose of the remainder of this chapter is to lay the foundation for future work in
solving these challenges.

130

7.4 Multiple Joint FASTA for Form Variation Inclusion


In order to confront the modeling challenges discussed in the previous section, a
proposed multiple joint FASTA process is outlined in this section. While the application
of this process is beyond the scope of this thesis, it is hoped that the presentation and
discussion of multilevel FASTA analysis will aid future work into this area.
To be able to model the complex case of form variation across multiple flexible
joints in an assembly, it is necessary to look at how each joint behaves separately. This
could be done by applying FASTA at each flexible joint separately (excluding the gap).
Since there is no rigid body misalignment at a flexible joint (mating surfaces are simply
placed together and then fastened), step 1 would be skipped. The second step should
proceed as usual, and includes the FEA modeling of flexible parts involved at each joint.
Step 3 of FASTA would also proceed as previously set forth, now requiring either surface
measurement data or surface continuity models for all flexible part mating edges. The
output from step 4 would be the mean and variance of the deformation and stresses seen
in each flexible part due to form variation at each joint.
With a set of statistical solutions from the application of FASTA at each flexible
joint, a slightly modified analysis of the gap would now be done to include the effects of
form variation. In addition to rigid contributions to misalignment, Step 1 of FASTA
would require the statistical inclusion of gap variation resulting from accumulated form
variation at flexible assembly joints. The FEA models would have already been created
during the analysis of flexible joints, leaving only the calculation of the equivalent
stiffness at the gap in Step 2. The third step of FASTA would proceed as usual, as would
131

the statistical FEA calculation (Step 4), though now including the adjusted gap
misalignment.
The steps for this multilevel FASTA process are outlined in Table 7-2, and
illustrated in Figure 7-3. It should be noted that this proposed method only accounts for
form variation along the line of fastening between mating surfaces. Perpendicular and
angular modes of variation propagation were not taken into account, but are discussed in
the next section.
Table 7-2 Outline of FASTA inclusion of form variation at flexible joints

Multiple Joint FASTA Application


1. Apply FASTA at Each Flexible Joint (excluding the gap)
Step 1 - Not needed since there is no initial gap
Step 2 - FEA of flexible parts at joint
Step 3 - Geometric covariance of flexible surfaces (requires data
or surface continuity model for each)
Step 4 - Statistical FEA solution
2. Modified FASTA Applied at Gap
Step 1 - Add statistically (RSS) the mean and variance of the rigid
gap variation and the mean and variance of the form
variation contribution from flexible joints.
Step 2 - FEA already done in analysis of flexible joints
Step 3 - Geometric covariance of gap flexible surfaces
Step 4 - Statistical FEA solution of modified gap variance

132

Multiple Joint FASTA Process


Part Geometry

Material Prop.
Surface
Continuity
Model

1. FEM

Part Variation

2. COV

[ ]

[K eq ]

Repeat for each


flexible joint

3. Statistical
FEA

( , )

FEA

FASTA Applied
at Flexible Joints
FASTA Applied at Gap

1. STA

{ i }

{ }, { }, { }
2
Ti

2
Ri

2. COV

2
SPi

[ ]

Surface
Continuity
Model

( , )

Stress( , )

4. Statistical
FEA
FEA

F( , )

Figure 7-3 Process flow diagram of multiple joint FASTA application

7.5 Modes of Variation Propagation


While the analysis discussed in the previous section may predict the effects of
form variation in flexible assemblies, it does so with perhaps excessive complexity. Even
133

with multiple levels of FASTA analysis, there are still degrees or modes of variation
propagation that may not be taken into account at the joints. Since some geometry in
flexible parts tends to act more rigidly than others, it may be possible to simplify the
modeling of flexible joints by categorizing some degrees of variation as essentially rigid.
This section looks at different modes of propagation with this perspective, using the
example assembly from Chapter 4 for illustration.
To begin, a rigid mode of variation propagation for flexible joints is defined as a
degree of variation propagation that acts essentially rigid due to geometry. A flexible
mode is then a degree of variation propagation in which some of the variation propagates
rigidly and some of the variation is absorbed by distortion of the compliant part(s) at or
near the joint. In order to illustrate this concept, the example assembly from Chapter 4 is
shown in Figure 7-4 with possible modes of variation propagation through flexible joints
labeled. Rigid propagation modes are noted with an R, and flexible modes with an F.
A

Gap

B
Ty

Ry

Rz

F
F

Tz

F
F

Flexible
Joint 2

Flexible
Joint 1

Figure 7-4 Variation modes at flexible joints for example assembly

134

Tx
Rx

In this figure, rotation about the x direction due to variation at flexible joint 1 is
labeled as a rigid mode. Looking at the geometry of part A, it is apparent that such
variation acts in the plane of the sheet of material. Since this is the stiffest direction for
the geometry, the assumption that variation propagates rigidly in this direction is
justifiable. Any distortion in the plane of the material will be negligible compared to outof-plane deflection. Variation at flexible joint 2 in Figure 7-4 has two rigid modes of
propagation: rotation about x and rotation about y. Again, both of these directions of
variation lie in the plane of the thin sheet of materialsuggesting that they will behave
rigidly. All other modes of variation propagation at the flexible joints in the example
assembly are considered flexible, with both rigid and elastic deflections possible.
To be able to predict how flexible modes of variation propagate, one must
determine what percent of the variation is absorbed in distortion and what percent is
transferred rigidly. Once quantified, the percent rigidity of a flexible mode could be used
to convert it to a rigid mode using some sort of a rigidity coefficient. Standard rigid joint
variation models could then be adapted to complete the analysis. Both flexible and rigid
will have a mean plus a variance. After closing the gap, the equilibrium profile will be
non-zero, in general. By adding the final profiles statistically, the mean and variance can
be determined.

7.6 Summary
As seen in the examples and discussion presented in this chapter, the treatment of
form variation propagation at flexible joints is complex. Parallels between rigid and
flexible joint variation propagation have been discussed and illustrated. Two proposed
135

methods for treating variation at flexible joints have been presented: a multiple level
FASTA analysis, and modeling using modes of propagation. It is hoped that the ideas
presented in this chapter will prove useful for future work in this area and the
incorporation of variation at flexible joints in the FASTA method.

136

CHAPTER 8

Application of FASTA Method


(Industry Example)

8.1 Introduction
In order to illustrate how the FASTA process is to be applied, the analysis setup
of an example assembly from industry is presented in this chapter. Although insufficient
data prevents an actual demonstration at the time of publishing, a detailed discussion of
the analysis steps and required inputs is given. The purpose of this chapter is to give an
example of the setup of FASTA on a real assembly, and to lay the groundwork for future
analysis and verification of the process on a real assembly. An assembly from industry is
first introduced, then each step of the FASTA analysis is setup and discussed.
Measurement of the assembly for analysis and verification is also discussed in detail.

8.2 Presentation of Assembly


To illustrate the application of the FASTA method in this chapter, a subassembly
of the leading edge flap of a Boeing 737 is used. The assembly contains 3 parts: an outer
aluminum skin, and two interior ribs. A fixture is used to hold the parts in place while
they are riveted together. Both the fixture and the parts come from the Boeing
Corporation, and are pictured in Figure 8-1.
137

a) Outer aluminum skin

b) Interior ribs

c) Assembly fixture shown with interior ribs in place


Figure 8-1 Pictures of parts and fixture for leading edge assembly

Figure 8-1c shows the interior ribs held in place by a spring-loaded member, with the
outer skin yet to be located by the outer shell of the fixture (seen at the top of the picture).
138

The holes marked by yellow on the fixture shell are used for the drilling and riveting of
the assembly along the curved surface of the outer skin. Additional riveting is done on
the back side of the fixture, along the flat portion of the outer skin.
This assembly is only about 12 inches in length, allowing for ease of
measurement. The design drawings available were originally made before 1990, and
only limited CAD data for these parts is available. The fixture was originally made
before 1980, but has had routine maintenance through the years. This assembly is an
excellent candidate for illustration of the FASTA process because it presents many of the
challenges involved in tolerance analysis.

8.3 Setup of FASTA


In this section each step of flexible assembly statistical tolerance analysis is
outlined and discussed using the leading edge assembly shown in Figure 8-1. Required
information that is missing from the available data is highlighted for each step and
suggestions are given to obtain the needed data.

8.3.1 Step 1: Determine misalignment


The first task in determining misalignment is to identify the assembly gap and
create a vector loop model that describes the gap at each fastening point. For the leading
edge assembly, the assembly gap occurs along the curved surfaces between the outer skin
and each interior rib part. The gap, at each of the five fastening points, is illustrated in
Figure 8-2 with mating surfaces labeled as A and B.

139

A
G5

G2

G3

G4

G1

Figure 8-2 Leading edge assembly gap at fastening nodes

A single generic or variable vector loop can be created to describe the gap at fastening
point i between surfaces A and B in terms of the fastening node locations. This vector
loop is illustrated in Figure 8-3, and is located by a datum reference frame defined by the
spar center-line (CL) and the flat base of the outer skin.
Magnified
View

n Ai

Ai

p Ai

Gi

pBi

nBi

Bi

X Ai

CL Spar

YBi
X Ai

YAi

YBi

X Bi

Figure 8-3 Generic vector loop describing the gap distance at node i

Of special note in the open vector loop presented in Figure 8-3, are the basic
dimensions (shown in boxes) used to describe the nominal locations of the fastening
140

nodes. Tolerances at each fastener location are described by position tolerances


referenced to the ideal location. This modeling technique is used because the fastening
locations are not specifically dimensioned. They are determined at assembly by the
drilling fixtures, and must be determined by measurement of the fixtures. By treating all
of the variation as zero-length vectors with normal and tangential variations, n and p at
each mating surface, the point locations can easily be inserted into the model using
Cartesian coordinates (once measurement data is available).
Vectors n Ai and n Bi describe variations normal to surfaces A and B, respectively,
at fastening location i. The orientations of these normal vectors are known from the CAD
model or by measurement, with constant angles Ai and Bi . Such variation could be
estimated from specified surface tolerance bands, or determined by measurement.
Vectors p Ai and p Bi are tangential to their respective surfaces, and describe positional
variation due to process variations in drilling and riveting.
With a generic vector loop that describes the gap at each fastening point, the next
task is to create the vector loop equations. For the x and y directions, the vector loop
equations are:
G xi = X Ai + X Bi + n Ai cos Ai + n Bi cos Bi + p Ai sin Ai + p Bi sin Bi

(8-1)

G yi = Y Ai YBi + n Ai sin Ai + n Bi sin Bi + p Ai cos Ai + p Bi cos Bi

(8-2)

The length of the mean gap vector Gi can be calculated using the Pythagorean Theorem
in terms of its x and y components. Since the nominal values of the normal and tangential
variation vectors are zero, the equation describing the mean gap length is simply:
141

[(

Gi = X Ai X Bi

) + (Y
2

Ai

YBi

)]
2

1
2

(8-3)

The next task is to write the partial derivatives of the vector loop equations in
terms of the variations. Because the basic dimensions and the normal angles are constant,
the partial derivatives of equations 8-1 and 8-2 include only the variations in the normal
and tangential vectors. These equations can be written symbolically as:
G xi =

G y i =

G xi
n Ai
G yi
n Ai

n Ai +

G xi
n Bi
G yi

n Ai +

n Bi

n Bi +

n Bi +

G xi

p Ai +

p Ai
G yi
p Ai

p Ai +

G xi
p Bi
G yi
p Bi

p Bi

(8-4)

p Bi

(8-5)

Evaluating the partial derivatives, equations 8-4 and 8-5 become:


G xi = n Ai cos Ai + n Bi cos Bi + p Ai sin Ai + p Bi sin Bi

(8-6)

G yi = n Ai sin Ai + n Bi sin Bi + p Ai cos Ai + p Bi cos Bi

(8-7)

While a matrix equation could be made to represent equations 8-6 and 8-7 at each of the
five mating points, it is just as easy to continue the analysis using the generic equations in
terms of fastening point i.
A statistical stack-up model can be applied to equations 8-6 and 8-7 to determine
the variation of the gapGi , at location i. The RSS of the variation components in the x
and y directions can be expressed as:

Gi = G x2i + G y2i

1
2

(8-8)

142

Substituting equations 8-6 and 8-7 into equation 8-8, with some algebra and
trigonometric simplifications, the gap variation at fastening point i can be written as:

{
} (

} (
(

n A2i + n B2i + n A2i + n B2i + 2 n Ai n Bi + p Ai p Bi cos Ai Bi

Gi = + 2 n Ai p Bi + n Bi p Ai sin Ai + Bi + 2n Ai p Ai sin 2 Ai

+ 2n Bi p Bi sin 2 Bi

1
2

(8-9)

Symbolic equations 8-3 and 8-9 statistically describes the misalignment between the
leading edge parts. Tangential surface variations can be treated separately from the
normal variation vectors, and would be obtained using surface scans along the mating
lines.
With gap mean and variation expressed symbolically, the challenge for the
completion of step 1 of FASTA is determining the location of the fastening points for the
leading edge assembly. While some of the riveting locations are specified in part and
tooling drawings, several key dimensions are missing. Additionally, common
manufacturing practices at the time the fixture was built put in doubt the correlation
between the designed riveting locations and the drilling guides on the actual fixture.
When the difference between ideal design and as-built tooling is significant,
measurement of the tooling for accurate fastening locations may be necessary. This is
especially the case when using analysis results for process verification. Measurement to
determine the fastening locations is discussed later in this chapter.

8.3.2 Step 2: Model compliant parts


Reviewing the leading edge assembly, it is apparent that all of the parts can be
treated as compliant since each one is made from a flat sheet of metal. With some
143

available CAD and drawing information, the ideal form of the parts can be used to create
FEA models. Existing CAD data includes solid models of the interior ribs, and a curved
surface that defines the outer shape of the skin. For this assembly CAD data is only
available in an old version of CATIA, requiring the use of IGES conversion for data
transfer. The shape of the outer skin can be reconstructed using the outer surface
definition from CAD and design dimensions from part drawings.
Close examination of the material used to form each part shows that the outer skin
is composed of a 2-layer aluminum laminate, while the interior ribs are formed using
only a single-layered aluminum sheet. In both cases the aluminum alloy used is 2024-T3,
with approximate material properties of: E = 10.501 Mpsi, and = 0.33.
Due to part and material geometry, element selection is an important step in
modeling the compliant parts of the leading edge assembly. In order to model the
laminate outer skin, a layered structural shell element could be used, such as SHELL99 in
ANSYS (see Figure 5-8). In addition to modeling the laminate nature of the outer skin,
SHELL99 also allows for offsetting of the shell element center line. This is a useful
feature when modeling geometry, given only a curve that defines the outer surface (not
the center-surface) of the part. Due to the rounded and curved geometry of the ribs, a
shell element that allows for significant curvature across a single element should be
selected when modeling these parts. SHELL93 in ANSYS matches this requirement.
As discussed in Chapter 5, meshing of the parts should be done such that nodes in
the model are located at each of the fastening points. Again, due to the nature of the asbuilt tooling, the placement of the mating boundary nodes is dependent upon
measurement data. Once obtained, such data points could be applied to the FEA model
144

using hard points, and then meshed. Preliminary FEA models of the leading edge parts
are shown in Figure 8-4.

a) FEA model of outer skin

b) Nominal ribs

Figure 8-4 Preliminary models of leading edge flexible parts, made in ANSYS

The mesh shown on the outer skin part is based upon CAD data and estimates, and must
be adjusted once accurate fastening locations are obtained through measurement. The
nominal geometry of the ribs is shown modeled in ANSYS, also awaiting required node
locations for final meshing.
Once meshed, the respective stiffness matrices of the flexible parts can be output
from ANSYS using sub-structuring analysis. Only relevant boundary degrees of
freedom at the fastening nodes would be included in these super-element matrices. In
order to proceed with step 2 of FASTA, only the fastening locations are lacking.

8.3.3 Step 3: Covariance calculation


In order to account for covariance, measurement data is again needed to describe
the surface characteristics of the mating part surfaces. Ideally, surface models for the
forming process would be available in the form of frequency spectra and could simply be
145

applied to determine geometric covariance using the techniques presented in Chapter 6.


Because these do not exist, surface measurement as presented by [Soman 1999] would
need to be done on multiple assemblies to represent the surface characteristics
statistically. Such measurement would need to be performed at each of the mating part
surfaces along the fastening line.
With measurement data available, the geometric covariance can be calculated
directly, without the need for a surface continuity model. Since the purpose of analyzing
this assembly includes verification of the analysis, surface measurement data could be
used to create a surface continuity model that could then be compared to the direct
covariance calculation. In this case, the modeling methods presented in [Tonks 2002]
should be used to create the needed models.

8.3.4 Step 4: Statistical FEA solution


Once solutions from the previous steps are obtained, the statistical force solution
can be easily calculated as discussed in Chapter 6. The mean displacements and stresses
throughout the interior would then be obtained by applying the mean gap displacements
as boundary conditions in the complete FEA model for each part and solving the load
system. A Covariance solution for displacement can be obtained through application of
Chapter 6 principles, though without automated methods it would be a challenge to sort
and keep track of the nodes in matrix manipulation. There is a need for software
assistance at this point of the analysis, to help automate the matrix manipulations and
keep track of node numbering referenced to part coordinate locations.

146

The covariance solution for stress also requires additional programming to incorporate
the statistical FEA outlined by [Bihlmaier 1999] into a manageable process for complex
geometry, such as seen in the leading edge assembly.
As is apparent from the requirements of this step, general software tools are
needed to take the principles and steps of FASTA into the realm of an efficient
application. Preliminary analysis software has been created, but now must be taken to the
next level for general use.

8.4 Measurement for Analysis & Verification


Application of the FASTA method to the leading edge assembly is dependent
upon the data acquisition outlined in the previous section. For this assembly,
measurement of the fixture to obtain accurate fastening locations is a first priority. Other
measurements must also be made for analysis verification purposes. Proposed methods
for achieving these measurements are presented and discussed in this section.

8.4.1 Measurement to determine fastening locations


In order to accurately determine the locations of the fastening points along the
curved mating surface of the leading edge assembly, measurement of the assembly
fixture is required. Among the many precise tools available in the field of metrology, a
coordinate measuring machine (CMM) holds the most promise for locating points on the
as-built fixture. This machine incorporates complex software and computer controlled
motion of a force sensitive probe to locate surface points in reference to desired

147

measurement datums. A picture of the assembly fixture on a Brown & Sharp CMM is
shown in Figure 8-5.

Figure 8-5 Assembly fixture positioned on CMM

By directing the CMM probe to contact multiple points inside of a single drilling
guide on the fixture, a cylindrical representation of that feature can be created in
reference to the assembly datums. The computer representation of the drilling guide can
be merged with CAD models of the parts, and used to find the intersecting centerline of
the drilling process on each part. This point of intersection represents the coordinate
location of a single fastening point. Repeating the process at each drilling guide provides
the location of each fastening point in the leading edge assembly.

148

8.4.2 Measurement for analysis verification


In order to verify the FASTA process, predicted statistical values must be
compared to measurement data from a population of assemblies. Due to inherent
difficulties in measuring forces and stresses in assembly, it is proposed that part
deformations should be used for analysis verification. By comparing measured part
displacements to predicted values from FASTA, the accuracy of the analysis method
could be quantified. In order to complete this type of verification, multiple
measurements of each part in a statistical set of assemblies are required.
For analysis verification using the leading edge assembly, surface scans of the
parts must be made before and after assembly. While a CMM can be used to obtain the
needed surface scans, care must be taken to assure that probe contact forces do not bias
the data by deflecting the compliant surfaces. Alternative metrology methods for surface
scanning should be researched. Possibilities include measurement using lasers and photo
imaging.

8.5 Summary
In this chapter the process of applying the FASTA method was illustrated using
an assembly from industry, and each step of the FASTA method was set up and
discussed. Required inputs were identified and suggestions were given for obtaining
these. Measurement methods for both analysis and verification of analysis were also
discussed. The groundwork for future analysis and verification using the leading edge
assembly has been set forth.

149

150

CHAPTER 9

Conclusions and
Recommendations

9.1 Introduction
The main objective of this thesis was to present a comprehensive, integrated
methodology for the statistical tolerance analysis of flexible assemblies. In order to do
this, rigid body tolerance analysis needed to be adapted for use with both rigid and
compliant parts, and incorporated into existing flexible part tolerance analysis methods.
For process integration, the methodology also needed to be summarized and presented in
a comprehensive manner for future application.
The objective and supporting aims of this thesis were achieved, and are
summarized in this concluding chapter. Strengths and limitations of the FASTA process
are listed, and suggestions for future work in this field of research are also presented.

9.2 Research Objectives


In order to present an integrated method for flexible statistical tolerance analysis,
or FASTA, four main tasks were completed. These are listed and summarized below:
1. Integration of rigid STA with FEA for variation analysis of assemblies with
compliant parts. To be able to treat complex assemblies that include both rigid and
flexible parts, the direct linearization method was first presented and adapted for
151

compliant application. Vector loop modeling using variable vectors was presented,
and rigid modes of assembly variation were identified and determined.

2. Use of commercial FEA software for modeling of compliant assemblies. The


needed inputs and specific steps were outlined for the modeling of compliant parts
using commercial FEA software. Sub-structuring analysis was presented as a means
for creating super-elements of entire parts and outputting reduced stiffness matrices
for linear force-deflection analysis. These methods were illustrated using ANSYS,
a well known FEA software package.

3. Inclusion of both compliant and rigid covariance components. In order to


statistically describe the behavior of flexible parts in assembly, the covariance of the
assembly system (at the mating gap) must be determined. This included the statistical
summation of both compliant surface continuity (geometric covariance), and rigid
gap covariance contributions. Zero and first order Legendre polynomials were
presented as a means to model the covariance of both rigid translational and rotational
terms.

4. Statistical FEA, incorporating rigid and compliant covariance. Utilizing the


covariance solutions at the gap, statistical solutions for interior deformations and
stresses needed to be shown. The covariant solution for interior deformations of
flexible parts was presented, and the groundwork for the covariant stress solution was

152

given. Groundwork for the statistical treatment of form variation at flexible joints
was also set forth.
Each of these main tasks was illustrated using an example assembly containing 1 rigid
and 2 flexible parts. The complete methodology was presented as a four step process,
with each step summarized and illustrated with a detailed process flow diagram. Finally,
an assembly from the aerospace industry was used to showcase the setup of the FASTA
process for analysis. A listing of the strengths and limitations of the FASTA method is
given in Table 9-1.
Table 9-1 Strengths and limitations of FASTA

FASTA Process Capabilities


Strengths
Limitations
Statistical solution of deformations, forces, and
FEA calculations, though
deformations with only 2 calls to FEA
efficient, are not trivial.
Inclusion of surface continuity and surface
waviness effects
Inclusion of both rigid and compliant part
variation sources

Complexity of model setup


Requires knowledge of
mating surface characteristics

Ability to model complex parts


Provides information on dimensional
contributions to variation, allowing for process
improvement
Adaptation to commercial CAD & FEA packages
Many of the limitations given in this table can be overcome by automation of the analysis
method in commercial software.

153

9.3 Thesis Contributions


Significant contributions made by the thesis in the field of statistical tolerance
analysis of flexible assemblies are:
Integration of rigid STA with FEA for variation analysis of assemblies with
compliant parts. A comprehensive system has been outlined in detail in four major
steps with required inputs and desired outputs. The system is suitable for use in the
design phase using specified tolerances, or in the production phase using measured
dimensional and surface variations. Previous research in statistical FEA, based on
material and geometric covariances, has been combined with rigid part dimensional
variation in a complete analysis system suitable for industrial applications.
The adaptation of the direct linearization method for use in determining
misalignment in flexible assemblies.
A separation of rigid body variation modes at flexible assembly gaps, identified as:
translational, rotational, and surface variation.
Including rigid covariance components in the gap covariance calculation, modeled
using Legendre polynomial transformation matrices. This made possible the
combining of rigid body, surface continuity, and elastic coupled variations into a
single covariance matrix for statistical FEA.
The use of commercially available FEA software for modeling complex flexible
assemblies and outputting reduced stiffness matrices using sub-structuring.
Detailed outline and diagram of the analysis process, which sets the groundwork
for software architecture.
154

The introductory treatment of form variation at flexible joints, including a proposed


multi-level FASTA process and the identification of rigid and flexible modes of
form variation propagation.
Demonstration of the complete process applied to a simple flexible assembly.
Example of analysis setup for an existing aerospace assembly, including the use of
vector loops with basic dimension and the resolution of variation to zero-length
vectors at fastening nodes.
Characterizing position variation and surface variation effects in statistical model of
gap

9.4 Suggested Future Work


While the FASTA process has been presented in detail, there are still facets of the
analysis that warrant more consideration and future study. A listing of suggested future
work is presented in this section that would help to further the analysis of flexible
assemblies.

1. Metrology and verification With the Steps of the FASTA analysis outlined in
detail and set up on an example assembly from industry, it is time to verify the
analysis. Future work in this area would include researching various metrology tools
for surface scanning, and comparing both predicted and measured part deformations.
This work would not only quantify analysis accuracy, but provide invaluable data for
process and analysis improvements.

155

2. Stress covariance calculation Although the basic theory is known for the
calculation of stress covariance, the inclusion of such in a commercial application
requires an efficient means for software implementation. Work in this area would
include a detailed look at the stress interactions between part elements, and
significant programming. It is suggested that software for the calculation of stress
covariance could be linked to conventional FEA software such as ANSYS.

3. Surface continuity modeling A method of characterizing the effects of specific


manufacturing processes on flexible part surfaces is needed in order to use surface
continuity models effectively. Research in this area would build upon the work done
by [Soman 1999], and would ideally produce a database of data for use in covariance
modeling based upon processes and part material. Additional work also needs to be
done to better characterize when to use Legendre versus Chebychev polynomial
series methods.

4. Form variation propagation Groundwork for the treatment of form variation


propagation at flexible joints has been presented in this thesis, but there is much left
to do. Research to determine how much variation is absorbed by flexible part
distortion and how much propagates rigidly and elastically through parts is needed.
Both simulation and measurement of real parts would be needed in this area of study.

5. Assembly order The inclusion of assembly order considerations in the FASTA


process would greatly increase its usefulness. This area of research could include
156

optimization of assembly order to minimize variation at critical dimensions. The


effects of tooling variation in the assembly process could also be included in this area
of research.

6. Inclusion of contact between fastening nodes In the FASTA process, force


closure of the assembly gap is modeled at the fastening nodes. A way to include
surface contact between mating nodes statistically could improve the accuracy of the
gap closure model.

7. Commercial software To be able to fully benefit from the analysis capabilities


proposed within this and previous works, the FASTA process must be incorporated
into commercial software. With preliminary analysis code and detailed process
architecture, this work has begun, but must now be taken to the next level of
application.

9.5 Closing Statement


The methods presented in this thesis provide a powerful new tool for the design
and analysis of flexible assemblies. With further advances and software implementation,
it holds great promise for lowering production costs in a multitude of industries.
Fixtureless assembly is being considered by many companies in the aerospace
industry. Before this can be achieved, a model for variation that accounts for all sources
of variation and predicts their accumulation and propagation in an assembly is needed.

157

The FASTA method presents such a model, allowing for: the prediction of variation
percent contribution, identification of variation sources that are most difficult to control,
and allowing for the focusing of process improvement and re-design for robust assembly.

158

APPENDIX
The appendix is divided into three sections as follows:
APPENDIX A: Vector Loop Modeling Rules and Regulations for 2-D
APPENDIX B: FEA Model Issues and Suggestions
APPENDIX C: Sub-structuring in ANSYS for Stiffness Matrix Output

159

160

APPENDIX A: Vector Loop Modeling Rules and Regulations for 2-D


Vector loop modeling rules presented by [Trego 1993] are presented below:
Table A-1: Vector loop modeling rules [Trego 1993]

I. Rules for Creating a Valid Set of Vector Loops


1. Loops must pass through every part and every joint in the assembly.
2. No single loop may pass through a part of joint more than once, but it may start
and end at the same point.
3. There must be enough loops to solve for all of the kinematic variables one loop
for every three variables.
II. Rules for Vector Paths
1. A vector loop must pass from one part to the next mating part through a common
joint.
2. The path across a part must pass though the Datum Reference Frame (DRF),
following the datum path from the incoming joint to the DRF and then following
the datum path from the outgoing joint to the DRF, only in reverse.
3. If the path across the part double back on itself with an equal and opposite vector,
the two vectors cancel each other and will be omitted from the loop.
III. Vector Loop / Joint Requirements (see figure below)
1. For a cylindrical slider joint, either the incoming or outgoing vector must be
normal to the sliding plane and end at the center of the cylinder.
2. For joints having a sliding plane (planar, cylindrical slider or edge slider), either
the incoming or outgoing vector must lie in the sliding plane.
3. For cylindrical joints (parallel cylinders in contact), the path through the joint
must start at the center of one cylinder and end at the center of the mating
cylinder, passing through the contact point.
IV. Rules for Open Loop Creation
1. The loop must pass through at least 2 parts and 1 joint
2. A single loop may not start and end at the same point.
3. There must be a minimum of one open loop for every design specification.
161

from
Datum 1

from
Datum 1

Datum 2

Datum 2

Edge Slider

Datum 1

Datum 2

Planar

Datum 1
R1
R1

R2

Datum 2

Parallel Cylinders

Cylindrical Slider
Figure A-1: Vector loop joint requirements

162

APPENDIX B: FEA Model Issues and Suggestions

Modeling Issues and Suggestions


Mastering, or even becoming familiar with FEA requires lots of practice. Even
when proficiency is obtained in one type of analysis application, new skills and tools may
be required for other analysis cases. This section in the appendix is dedicated to
modeling issues and suggestions intended to help the reader better understand the
intricacies of modeling in FEA for application in FASTA. Topics discussed include:
limiting degrees of freedom, node locating using hardpoints, mesh accuracy,
bookkeeping, file conversion from CAD, and automation using batch files.

Limiting Degrees of Freedom (DOF)


In FEA, stiffness matrix size is a significant factor in the computational time
required to obtain a solution. Without paying attention to this fact it is easy to create a
model that needlessly uses valuable analysis time. An example of how easily stiffness
matrix size can get out of hand may be illustrated using a 10 inch square plate. If square
elements with a 1 inch edge-length are used to mesh the plate, then there are a total of 11
nodes along each edge and a total of 121 nodes on the surface (Figure B-1).

163

10

1
1

10

Figure B-1: Example of DOF for a simple plate

With displacement in all 6 degrees of freedom under consideration at each node, the total
number of DOF for the plate is 6 121 or 721. This means that the stiffness matrix for
this case has 721 rows and 721 columns to make a total of 527,076 matrix elements. If
the element size is reduced to 0.5 inches then there are a total of 441 nodes, 2646 total
DOF and 7,001,316 matrix elements. Even fast computer processors take a while to
perform operations on such large matrices.
Another way to significantly reduce stiffness matrix size and computation time in
FEA is to limit the number of degrees of freedom considered in the analysis. A typical
3D element allows up to six degrees of freedom at each node (illustrated in Figure B-2).

Ty

Ry

Rz

Tx
Rx

Tz

Figure B-2: Six degrees of freedom at one node

For the relatively small deflections seen in FASTA, many of these degrees of freedom
have negligible displacement and can be omitted. For the example assembly only the y164

direction was considered, reducing the total number of degrees of freedom per part to 5.
As is always true when making simplifying assumptions, care should be taken to verify
that the omitted DOF are truly negligible in the analysis being done.

Node location (hard points)


As discussed in Chapter 5, it is important that nodes are places at each of the
fastening points under consideration in FASTA. However, as part complexity increases
so does the difficulty of manually controlling node location. The solution to this
challenge is the use of hard points, set locations where automated meshing is required to
place a node. In ANSYS, hard points can be easily placed on a surface using global
coordinates. With the use of hard points, however, automated meshing is limited to a free
mesh.

Accuracy of mesh (refining)


As suggested by its name, finite element analysis simplifies the analysis of parts
under complex loading conditions by dividing the problem into discrete portions, called
elements, which can be solved more easily. The accuracy of the analysis is then
dependant upon how finely divide or meshed the part is. However, more element
divisions also results in greater computational time, and a point of limiting returns is
often reached. In order to achieve the desired analysis accuracy with the minimum
amount of time, it is important to plan part meshing accordingly. For example, locations
of stress risers, such as holes and fillets in a structure, are more critical for stress concerns
and thus should be more finely meshed than surrounding areas. A simple way to check if
the mesh density is sufficient for the analysis being done is to solve the load case, then
165

double the double the number of elements and repeat the analysis. If the results for both
cases are nearly identical, then the original mesh density was probably sufficient.
ANSYS has many tools for automating mesh refining, whether for a whole part or just a
particular area. For more information on this topic, see the ANSYS users manuals.

Bookkeeping
One of the challenges in using data output from finite element analysis is keeping
track of the node numbers and their corresponding part locations. When using reduced
stiffness matrices output from commercial FEA software, it is important to make sure that
subsequent operations apply loading conditions to the correct nodes within each matrix.
For example, stiffness matrices output from ANSYS using sub-structuring methods (see
appendix C), order the matrix indices by ascending node number rather than by location.
For this reason, bookkeeping becomes extremely important when trying to keep track of
node locations in FASTA analysis. Some suggestions for good bookkeeping strategies
include: using tables, figures, and taking care when sorting data externally from the FEA
model. As careful bookkeeping is practiced, mistakes and analysis errors will be much
more easily avoided.

CAD conversion (IGES)


Though much improved over the years, cross-package data transfer between CAD
systems and analysis software is nowhere near perfect. Problems that can occur in data
transfer from CAD to FEA software include: discontinuities and connectivitybreakdowns between geometry. Many FEA packages, including ANSYS, contain
software routines that search out and fix such errors, though not always perfectly. If
166

geometry simply cannot be transferred from a CAD file into FEA software, a separate
FEA model may need to be constructed.

Batch files automation


Many sophisticated FEA software packages have the ability to read in specific
coding languages for analysis automation. In ANSYS such files are called batch files,
and are generally appropriate only for simple geometry. The possibility of automating
the sub-structuring process using batch files is worth looking into when parametric
models are required for repeated analysis. For more information on the use of batch files,
see the ANSYS tutorials and users manuals.

167

168

APPENDIX C: Sub-structuring in ANSYS for Stiffness Matrix Output

Introduction
As discussed in Chapter 5, sub-structuring analysis in ANSYS provides an
automated means for obtaining reduced stiffness matrices of flexible parts with complex
geometry. In this section of the appendix, a step-by-step tutorial is given for performing
this analysis in ANSYS version 5.7 on a UNIX machine. Each step is briefly discussed
and includes a graphical menu path sequence if applicable.

Sub-structuring Analysis in ANSYS


Step 1: Create and mesh part geometry
In order to perform sub-structuring analysis, a part model must first be created
and meshed. Care should be taken to place nodes at the boundary locations to be
used for the reduced matrix analysis. For FASTA application, these boundary
nodes should correspond to fastening point locations at the gap being analyzed.
Step 2: Specify the analysis type
Once the model has been created, sub-structuring analysis is selected using the
following menu path:
Solution>New Analysis>Substructuring>OK
Step 3: Select option to print stiffness matrix
In order to output the stiffness matrix, the default must first be changed.
169

Solution>Analysis Options
With the sub-structuring analysis options window open, enter a file name to be
assigned to the super-element matrix file. Make sure that the matrix to be
generated is: Stiffness, and select the items to be printed as: LoadVect+Matrix.
Then hit OK.
Step 4: Choose master nodes and master dofs
In order to create the super-element matrix, the boundary nodes must be chose. In
ANSYS these are called master nodes. Follow the menu path below and choose
the master nodes either graphically or by number.
Solution>Master DOFs>Define>select master nodes>OK
Once the master nodes are chosen, the model can be further reduced by limiting
the degrees of freedom included in the stiffness matrix. Select the desired DOFs
and hit OK.
Step 5: Apply desired boundary conditions
If desired, boundary constraints can be applied to the model at this time. Keep in
mind that any constraints at master nodes will override the selected degree of
freedom, and will not be output in the stiffness matrix.
Step 6: Solve the current load system
With master degrees of freedom defined, the system can be solved.
Solution>solve current LS>OK
Step 7: Retrieve the stiffness matrix output

170

Once the analysis is complete, the matrix output can be retrieved from the superelement matrix file created. This is done using the top pull-down control menu
with the following selections:
List>Other>Superelem>Data
A window entitled List Superelement Data appears. In the box labeled: Name
or no. of superelem, enter the file name chosen in step 3 to represent the matrix.
Under data to be listed, select Full contents, and hit OK. A window now appears
that include information about the super-element, as well as the stiffness matrix.

The stiffness matrix is listed by row, and is read from left to right at each line.
Above first row of the stiffness matrix, a list of the master node numbers and
corresponding degrees of freedom is given in numerical order. This is also the
organizational order of the stiffness matrix, the first node corresponding to the
first row and column, and so on. This is an important bookkeeping tool, and
should be saved. The file can be saved using File>Save As>etc., and then parsed
by hand or using a simple C-program. The reduced part stiffness matrix can now
be used for further steps in the FASTA process.

171

172

REFERENCES
ASME, 1994, Dimensioning and Tolerancing, ANSI Y14.5M-1994, ASME, New
York, NY.

Bihlmaier, B., 1999, Tolerance Analysis of Flexible Assemblies Using Finite Element and
Spectral Analysis, M.S. Thesis, Brigham Young University, Provo, Utah.

Chang, M., Gossard, D.C., 1997, Modeling the Assembly of Compliant, Non-ideal
parts, Computer-Aided Design, Vol. 29, No. 10, pp. 701-708.

Chase, K.W., Gao, J., Magleby, S.P., Sorensen, C.D., 1996, Including Geometric
Feature Variations in Tolerance Analysis of Mechanical Assemblies, IIE Transactions
(Institue of Industrial Engineers), Vol. 28, No. 10, Oct. 1996, pp.795-807.

Chase, K.W., Magleby, S. P., Gao, J., 1997, "Tolerance Analysis of 2-D and 3-D
Mechanical Assemblies with Small Kinematic Adjustments," Advanced Tolerancing
Techniques, JohnWiley.

173

Chase, K.W., Parkinson, A.R., 1991, A Survey of Research in the Application of


Tolerance Analysis to the Design of Mechanical Assemblies, Research in Engineering
Design, 3:23-37.

Crisfield, M.A., 1986, Finite Elements and Solution Procedures of Structural Analysis,
Swansea, UK: Pineridge Press Limited.

Gao, J., Chase, K.W., Magleby, S.P., 1998a, Global Coordinate Method for Determining
Sensitivity in Assembly Tolerance Analysis, Proceedings of the ASME International
mechanical Engineering Conference and Exposition, Anaheim, California.

Gao, J., Chase, K.W., Magleby, S.P., 1998b, Generalized 3-D Tolerance Analysis of
Mechanical Assemblies with Small Kinematic Adjustments, IIE Transactions (Institue
of Industrial Engineers), Vol. 30, No. 4, April 1998, pp.367-377.

Hockmuth, R., Meerkamm, H., Schweiger, W., 1998, An Approach to a General View
on Tolerances in Mechanical Engineering, 2nd International Workshop on Integrated
Product Development IPD 98; Magdeburg, pp. 65 - 76.

Hu, M., Lin, Z., Lai, X., Ni, J., 2001, Simulation and Analysis of Assembly Processes
Considering Compliant, Non-ideal Parts and Tooling Variation, International Journal of
Machine Tools and Manufacture, Vol. 41, No. 15, pp. 2233-2243.

174

Johnson, R.A., Wichern, D.W., 1988, Applied Multivariate Statistical Analysis, Prentice
Hall: New Jersey.

Liu, S.C., Hu, S.J., 1995, An Offset Finite Element Model and its Application in
Predicting Sheet Metal Assembly Variation, Intl. Journal of Machine Tools and
Manufacture, Vol. 35, No.11, pp. 1545-1557.

Liu, S.C., Hu, S.J., Woo, T.C., 1996, Tolerance Analysis for Sheet Metal Assembly,
Journal of Mechanical Design, Transactions of the ASME, Vol. 118, pp. 62-67.

Liu, S.C., Hu, S.J., 1997, Variation Simulation for Deformable Sheet Metal Assemblies
Using Finite Element Methods, Journal of Manufacturing Science and Engineering,
Transactions of the ASME, Vol. 119, pp. 368-374.

Merkley, K., Chase, K.W., Perry, E., 1996, An Introduction to Tolerance Analysis of
Flexible Systems, MSC World Users Conference.

Merkley, K., 1998, Tolerance Analysis of Compliant Assemblies, Ph.D. Dissertation,


Brigham Young University, Provo, Utah.

Narahari, Y., Sudarsan, R., Lious, K.W., Duffey, M.R., Sriram, R.D., 1999, Design for
Tolerance of Electro-Mechanical Assemblies: An Integrated Approach, IEEE
Transactions on Robotics and Automation, Vol. 15, No. 6, pp. 10662-1079.
175

Sellem, E., Rivire, A., 1998, Tolerance Analysis of Deformable Assemblies,


Proceedings of DETC98: 1998 ASME Design Engineering Technical Conference,
Atlanta, GA., 7 pages.

Sellem, E., Rivire, A., De Hillerin, C.A., Clement, A., 1999, Validation of the
Tolerance Analysis of Compliant Assemblies, Proceedings of DETC99: 1999 ASME
Design Engineering Technical Conference, Las Vegas, Nevada, 6 pages.

Shiu, B.W., Ceglarek, D., Shi, J., 1997, Flexible Beam-Based Modeling of Sheet metal
Assembly for Dimensional Control, Transactions of NAMRI/SME, Vol. 25, pp. 49-54.

Soman, S., 1999, Functional Surface Characterization for Tolerance Analysis of Flexible
Assemblies, M.S. Thesis, Brigham Young University, Provo, Utah.

Stout, J., 2000, Geometric Covariance in Compliant Assembly Tolerance Analysis, M.S.
Thesis, Brigham Young University, Provo, Utah.

Trego, A., 1993, A Comprehensive System for Modeling Variation in Mechanical


Assemblies, M.S. Thesis, Brigham Young University, Provo, Utah.

Tonks, M., 2002, A Robust Geometric Covariance Method for Flexible Assembly
Tolerance Analysis, M.S. Thesis, Brigham Young University, Provo, Utah.
176

Anda mungkin juga menyukai