Вы находитесь на странице: 1из 83

UNIVERSITY OF CINCINNATI

August 30th
_____________ 03
, 20 _____

Rami A. Musa
I,______________________________________________,
hereby submit this as part of the requirements for the
degree of:

Master's of Science
________________________________________________
in:
Industrial Engineering
________________________________________________
It is entitled:
Simulation-Based Tolerance Stackup
________________________________________________
Analysis for Machining
________________________________________________
________________________________________________
________________________________________________

Approved by:
________________________
Dr. Samuel Huang (Chair)
________________________
Dr. Sam Anand
________________________
Dr. Richard Shell
________________________
________________________
Simulation-Based Tolerance Stackup Analysis in Machining

A thesis draft submitted to the

Division of Research and Advanced Studies

of the University of Cincinnati

in partial fulfillment of the

requirements for the degree of

MASTER OF SCIENCE

In Industrial Engineering

in the department of Mechanical, Industrial and Nuclear Engineering

of the College of Engineering

August, 2003

by

Rami A. Musa

Bachelor of Science in Mechanical Engineering

Jordan University of Science and Technology, 1999

Committee Chair: Dr. Samuel H. Huang


Abstract

Dimensional and geometric tolerance can result from either process variation and/or process

stackup tolerance. Tolerance stackup (accumulation) is an important topic in machining that

is interrelated with tolerance control, tolerance allocation and setup planning. During

machining operations of a part, tolerance stackup is inevitable most of the time. Therefore,

tolerance stackup must be studied accurately and efficiently. In spite of this, traditional

methods for analyzing stackup (statistical and worst-case methods) have some drawbacks that

reduce their accuracies. These drawbacks are discussed in details. This study presents a novel

method for analyzing tolerance stackup in three dimensional-space by simulating machining

and inspection process using Monte Carlo simulation along with major manufacturing errors.

It overcomes the argued drawbacks in the traditional methods. Further, it is proved that both

the statistical and worst-case methods are conservative compared to the proposed one.

Therefore, simulation-based tolerance stackup analysis is more cost-effective as it gives more

chances to accept process plans that are usually precluded using the traditional ones. Three

illustrative examples are presented to compare the results of the simulation with the

traditional methods.

I
II
Acknowledgments

First of all, I wish to offer my sincerest gratitude to my advisor; Dr. Samuel Huang who was

an outstanding advisor in all measures during my work with him. His professionalism,

knowledge and keenness inspired and taught me a lot.

True thanks to Dr. Sam Anand and Dr. Richard Shell for serving as committee members in

my thesis defense, words of encouragement and appraising my effort. Also, I would love to

thank and recognize my friends: Mohammad Hamdan, Mohammad Younis and Zain Dewaik,

who introduced and encouraged me all the way to go for my graduate study. Also, I would

like to extend my thanks to my colleague and friend Anshum Jain who contributed

significantly in conducting the experiment. Most prominently, my deepest gratefulness is to

my family for their encouragement and support. I always felt I am the luckiest person in the

world to have such a family; my late father, my loving mother and my brothers: Naji and

Husam.

This work has been gratefully sponsored by the National Science Foundation and thankfully

collaborated with Delphi Automotive Systems in Dayton, Ohio.

III
Contents

1. INTRODUCTION 6

1.1 BACKGROUND AND MOTIVATION 6

1.2 OBJECTIVES OF THE RESEARCH 12

1.3 THESIS ORGANIZATION 13

2. LITERATURE REVIEW 14

2.1 BASIC CONCEPTS 14

2.2 TOLERANCE STACKUP; DEFINITION AND APPLICATIONS 18

2.3 TRADITIONAL ANALYTICAL TOLERANCE STACKUP ANALYSES 20

2.3.1 WORST-CASE ANALYSIS 22

2.3.2 STATISTICAL ANALYSIS 22

2.4 TOLERANCE CHART 23

3. SIMULATION-BASED TOLERANCE STACKUP ANALYSIS 25

3.1 SIMULATION ARCHITECTURE 25

3.1.1 MONTE CARLO SIMULATION 29

3.1.3 MANUFACTURING ERRORS 30

3.1.3.1 ERROR CATEGORIES 30

3.1.3.2 MACHINING ERROR (CUTTING TOOL DEVIATION) 34

3.1.3.3 LOCATING/CLAMPING DEVIATION (FIXTURE UNIT ERROR) 35

3.1.3.4 RAW PART ERROR 35

3.1.5 ERROR SYNTHESIS (AGGREGATION) 35

3.2 VIRTUAL INSPECTION 36

1
3.2.1 DATUM EVALUATION 36

3.2.2 EVALUATION ALGORITHMS FOR GD&T 41

3.3 SAMPLE PLAN 43

3.4 STOPPING (TERMINATING) CRITERIA 45

4. MANUFACTURING ERROR EVALUATION 48

4.1 MACHINING ERROR EVALUATION 48

4.2 LOCATING/CLAMPING ERROR (FIXTURE UNIT ERROR) EVALUATION 51

4.3 RAW PART ERROR EVALUATION ALGORITHM 54

5. ILLUSTRATIVE EXAMPLES 58

5.1. EXAMPLE 1: TWO MACHINING OPERATIONS (WITHIN ONE SETUP) 58

5.2. EXAMPLE 2: FOUR MACHINING OPERATIONS (IN THREE SETUPS) 60

5.3. EXAMPLE 3: ABS PART 61

6. CONCLUDING REMARKS AND RECOMMENDATIONS 68

6.1 SUMMARY 68

6.2 RECOMMENDATIONS FOR FUTURE WORKS 69

BIBLIOGRAPHY 71

2
List of Figures

Figure 1-1: One way clutch mechanism [Chase, Gao and Magleby (1994)].............................8

Figure 1-2: One way clutch mechanism vector loop [Chase, Gao and Magleby (1994)] .........8

Figure 1-3: 1-D assembly mechanism [Law (1995)]...............................................................10

Figure 1-4: 2-D Closed vector loop for one way clutch mechanism [Chase, Gao and Magleby

(1991)]..............................................................................................................................10

Figure 1-5: 3-D Closed vector loop for crank slider mechanism [Chase, Gao and Magleby

(1991)]..............................................................................................................................10

Figure 1-6: Ideal process condition..........................................................................................13

Figure 2-1: With Cp=1, only 2700 part per million (PPM) defects are expected ....................17

Figure 2-2: Mean drift in processes .........................................................................................17

Figure 2-3: Typical setup planning approach ..........................................................................20

Figure 2-4: Dimension Chain of c, 2 links, 1D........................................................................21

Figure 2-5: Dimension Chain of c, 4 links, 1D........................................................................21

Figure 2-6: Example of tolerance chart [Xue and Ji (2002)] ...................................................24

Figure 3-1: Part representation by sample points ....................................................................26

Figure 3-2: System Architecture..............................................................................................26

Figure 3-3: Monte Carlo Simulation (source:

http://www.ymp.gov/documents/ser_b/figures/chap4_2/f04-174.htm)...........................30

Figure 3-4: Error models used in this study.............................................................................33

Figure 3-5: Setup Error ............................................................................................................34

Figure 3-6: Combined effects of setup and machining errors..................................................34

Figure 3-7: Translated least-squares approach ........................................................................37

Figure 3-8: Candidate datum set approach ..............................................................................39

Figure 3-9: 2-D projection of the convex hull [Wilhelm (1998)]............................................40

3
Figure 3-10: Non-rejected datum [Wilhelm (1998)] ...............................................................40

Figure 3-11: Sample points located in a square feature using different approaches ...............45

Figure 3-12: Sample point locations using random and low-discrepancy methods [Davis and

Martin (1998)]..................................................................................................................45

Figure 3-13: Benchmarked results at 1 billion iterations [Cvetko, Chase and Magleby (1998)]

..........................................................................................................................................47

Figure 4-1: Dial Indicator measuring machined surface..........................................................50

Figure 4-2: Fixture Unit with the workpiece ...........................................................................52

Figure 4-3: Part Surfaces .........................................................................................................53

Figure 4-4: Coordinate measuring machine (CMM) ...............................................................56

Figure 4-5: Fixture unit with CMM probe...............................................................................56

Figure 5-1: Example 1 (design requirements and the machining sequence) ...........................58

Figure 5-2: Example 1 results ..................................................................................................60

Figure 5-3: Example 2 (design requirements...........................................................................61

Figure 5-4: Example 2 output (The simulation output for the distributions of the distances

between surfaces).............................................................................................................61

Figure 5-5: Example 3; ABS (Antiblock System) housing, Bosch (Source:

http://www.wzl.rwth-aachen.de/WM/SIMON/deliverables/DA0/DA0_02D.htm) .........62

Figure 5-6: ABS dimensional requirements ............................................................................62

Figure 5-7: ABS part setup plan ..............................................................................................64

Figure 5-8: Tolerance Chart of ABS part ................................................................................65

Figure 5-9: Dimensions histogram using simulation ...............................................................66

Figure 5-10: Progress of results with sample size increase .....................................................66

Figure 5-11: Rejection areas comparison when allocating concluding links tolerance using

worst case, statistical and simulation methods ................................................................67

4
List of Tables

Table 2-1: Geometric Tolerances (ASME Y14.5M-1994) ......................................................16

Table 3-1: Manufacturing Error Classification........................................................................31

Table 3-2: Manufacturing Error Models..................................................................................32

Table 3-3: Recommended sample size for different geometries [Henzold (1995)].................44

Table 4-1: Data Collection.......................................................................................................49

Table 4-2: Machining Error Data.............................................................................................50

Table 4-3: Data Collection.......................................................................................................54

Table 4-4: Variance comparison between simulation and experiment for smooth part ..........57

Table 4-5: Variance comparison between simulation and experiment for a rough part..........57

Table 5-1: Tolerance stackup evaluation comparison for example 1 ......................................60

Table 5-2: Tolerance analysis results for example 2 ...............................................................61

Table 5-3: Simulation results at 100,000 iterations .................................................................63

Table 5-4: Tolerance evaluation using the three approaches...................................................64

Table 5-5: Part per million (PPM) rejections comparison when allocating tolerance using

worst case, statistical and simulation methods ................................................................67

5
1. Introduction

1.1 Background and Motivation

Tolerance is a common arguing point between design and manufacturing. Design engineer

tends to tighten the tolerance to meet functional requirements whereas production engineer

tends to loosen (relax) it to satisfy resource availability. Nevertheless, the most important

factor to be considered is the cost. Cost increases hysterically by tightening the tolerance.

However, since tolerance is inevitable as it is impossible to have perfectly accurate

machining, raw part, fixture unit and measurement machine, it has to be compromised by

different departments in the companies. Obviously, tolerance problem is kind of promoter for

concurrent engineering work among organization departments; namely: design,

manufacturing, customer service and management.

One serious problem in process planning is that some good plans (plans that lead to design

requirement satisfaction) could be rejected and some bad plans could be accepted due to

inaccurate traditional methods of evaluating tolerance stackup. Tolerance stackup can be

defined as the accumulation (or stackup) of errors when machining a part using different

operational datum than the ones specified in the blueprints. The two traditional methods used

nowadays to analyze tolerance stackup in machining are: worst-case and statistical methods.

These methods are believed to have major drawbacks that reduce the accuracy of tolerance

stackup evaluation. These drawbacks are:

1. Worst-case is exaggeratedly pessimistic in calculating tolerance stackup.

2. Statistical analysis assumes independency between dimensions. Further, statistical

analysis assumes that the contributing links are normally distributed.

6
3. Tolerance stack between features is preformed in one dimension; which does not

represent the actual three-dimensional features of interest. 3-D simulation must be the

driving force behind the entire dimensional management process [Craig (1996)].

4. Manufacturing errors are not taken into account.

5. Geometric tolerance stackup cannot be estimated. The stackup of geometric tolerance

was usually ignored [Lin and Zhang (2002)].

Additionally, we found that both of the traditional methods evaluate tolerance stackup

conservatively. In this work we developed a more accurate method for evaluating tolerance

stackup in machining that can lead to more cost-effective (less conservative) and/or less

tighter plans. Our method overcomes the above-mentioned drawbacks by simulating

machining and inspection processes along with major manufacturing errors using Monte

Carlo simulation. It will be shown in chapter 5 (illustrative example 3) that using our method

for stackup evaluation will result in much less rejects expectations per million parts compared

to the traditional methods using the same resources.

Machining Tolerance Stackup vs. Assembly Tolerance Stackup

Some research works have been done in assembly tolerance stackup using Monte Carlo

simulation in the literature. Although this seems quite close to our work here in tolerance

stackup for machining, there are exclusive differences between the two problems, their

formulations and applications. Component (part) and assembly designs are the two major

tasks in any design department. Component design provides a single component drawing that

include dimensions; and dimensional and geometric tolerances. Some examples of

components are: shaft, gear, pulley, etc. However, it is unlikely to have a component

functioning alone as there is a need to assemble it with other components. Assembly design

7
studies the suitability of two or more components to meet machine functions. When

assembling parts together, there should be some manufacturing variation in the part that will

cause assembly tolerance stackup. Figures 1-1 and 1-2 show an example of an assembled one

way-clutch mechanism. The mechanism consists of: four rollers, a hub, four springs and an

outer ring. The objective of the tolerance analysis here is to study the effect of manufacturing

errors in component dimensions (a, e, c) on assembly dependent dimensions ( Φ1 , b ).

Figure 1-2: One way clutch


Figure 1-1: One way clutch mechanism
mechanism vector loop [Chase, Gao
[Chase, Gao and Magleby (1994)]
and Magleby (1994)]

This problem has been studied extensively by Chase, Magleby and Gao in Brigham Young

University. They developed computer software (CATS) that applies their methods in

assembly tolerance analysis. Another system has been developed by Variation System

Analysis (VSA).

The first step in evaluating assembly tolerance stackup in the literature is to find an explicit

function of the dimension (tolerance) to be controlled in terms of the other components using

8
trigonometric functions. The following are the explicit functions of the assembly dimensions

for the mechanism shown in figure 1-2 [Gao, Chase and Magleby (1997)]:

a+c
Φ1 = cos −1 ( ) (1.1)
e−c

b = (e − c ) 2 + ( a + c ) 2 (1.2)

Some assembly tolerance stackup methods in the literature that assumes the availability of

explicit assembly functions are [Gao, Chase and Magleby (1997)]:

1. Linearization of the assembly function using Taylor series expansion,

2. Method of system moments,

3. Quadrature,

4. Monte Carlo simulation,

5. Reliability index,

6. Taguchi method.

Normally, it is very hard or even impossible to get explicit assembly equations for a typical

assembly mechanism. Vector-loop-based assembly models use vectors to represent the

dimensions in an assembly that can be used to find a set of implicit assembly equations.

Figures 1-3, 1-5 and 1-6 show examples of closed vector loops for 1-D, 2-D and 3-D

mechanisms.

9
Figure 1-3: 1-D assembly Figure 1-4: 2-D Closed vector Figure 1-5: 3-D Closed vector
mechanism [Law (1995)] loop for one way clutch loop for crank slider mechanism
mechanism [Chase, Gao and [Chase, Gao and Magleby (1991)]
Magleby (1991)]

The following are the governing assembly equations for the closed loop one-way-clutch

shown in figure 1-2 [Gao, Chase and Magleby (1997)]:

hx = 0 = b + c sin(Φ1 ) − e sin(Φ1 )
h y = 0 = a + c + c cos(Φ1 ) − e cos(Φ1 ) (1.3)
hθ = 0 = 90 − 90 + 90 − Φ1 − 180 + Φ 2 + 90 = −Φ1 + Φ 2

From the third equation in the previous set of equations (1.3), it can be seen that Φ1 = Φ 2 = Φ .

This reduces the equations into two as follows [Gao, Chase and Magleby (1997)]:

hx = 0 = b + c sin(Φ ) − e sin(Φ )
(1.4)
h y = 0 = a + c + c cos(Φ ) − e cos(Φ )

It is apparent that it is very difficult to convert these equations into explicit form. The main

two methods available in the literature to solve this problem for implicit assembly functions

are: Direct Linearization Method (DLM) and Monte Carlo Simulation. First order Taylor

10
series linearizes the assembly constraints in DLM to have a set of linear simultaneous

equations and then linear algebra is used to solve them. Afterwards, assembly tolerance

stackup are estimated using statistical or worst case methods. Monte Carlo simulation method

includes the following steps: (1) generate random variates for each variable in the assembly

constraints (2) Select appropriate nonlinear solvers to solve the constraints (3) fit the output

numbers with a distribution and get its parameters (first four moments: mean, variance,

skewness and kurtosis.) Chase, Gao and Magleby use Crystal Ball software to solve the

problem. Crystal Ball is spreadsheet Monte Carlo Simulation software that can solve implicit

nonlinear simultaneous equations.

Gao, Chase and Magleby (1995) made a comparison between the two methods. It turned out

that the concern regarding the DLM is the accuracy and the concern regarding the Monte

Carlo simulation is the huge number of iterations needed to solve the problem.

Noteworthy, the following are the differences between using Monte Carlo simulation for

machining tolerance stackup analysis [Musa and Huang (2003)] and assembly tolerance

stackup analysis [Chase, Gao and Magleby]:

(1) Method. Machining tolerance stackup analysis simulates manufacturing variations

whereas assembly tolerance analysis simulates component variations because of

manufacturing variations. The two analyses are close in the sense that we are trying to

maintain a tolerance for a critical component in the case of assembly tolerance

analysis and a concluding link in the case of Part tolerance analysis.

(2) Objective. The objective of assembly tolerance analysis is to assign tolerances for all

the assembly components to maintain a specific tolerance for a critical component

whereas the objective of machining tolerance stackup analysis is to evaluate the

11
goodness of a process plan and/or assign proper contributing link tolerances

(increasing and decreasing links) to maintain concluding link tolerance.

(3) Sequence. Assembly tolerance analysis comes after machining tolerance stackup

analysis.

(4) Independence. It is safe to say that mechanical components’ variations are

independent which is not the case for machined features in machining.

1.2 Objectives of the Research

Improving quality and reducing cycle time and cost are the main objectives for competitive

manufacturing these days. In other words, achieving minimum tolerance possible using the

available resources, reducing trial and error procedures and taking economical issues into

consideration can lead to the ideal process which all industries aim at (figure 1-6). These

objectives can be achieved partially by effectively controlling the tolerance in manufacturing.

Tolerance control involves controlling the tolerance stackup via proper choices of processes,

process sequence, and locating datums.

The objective of this study is to present a novel, less conservative and more accurate

evaluation method of tolerance stackup compared to the existed analytical ones (worst case

and statistical methods) in the literature. This method is based on simulating machining and

inspection process using Monte Carlo simulation along with major manufacturing errors.

12
IDEAL PROCESS

ITERATIVE
TIGHTEST
DESIGN/ ECONOMICALLY
TOLERANCE
MANUFACTURE FEASIBLE
POSSIBLE
AVOIDANCE

Figure 1-6: Ideal process condition

1.3 Thesis Organization

The thesis is divided into six chapters. It starts in chapter 1 with the introduction that explores

background of the problem, motivation and objectives of the work. Then, chapter 2 reviews

some basic concepts and terms that are commonly used in later chapters and discusses the

problem of tolerance stackup by defining it, presents the traditional analytical methods

available in the literature and discusses tolerance chart method. Monte Carlo simulation

based tolerance stackup method is described in chapter 3; in which simulation architecture,

manufacturing error categories and models, sample plan and stopping criteria for the

simulation are illustrated. Afterwards, experiment procedures, requirements and algorithms

for evaluating: machining, fixture unit and raw part errors are outlined in chapter 4. In

chapter 5, three illustrative examples are demonstrated and solved using the proposed method

and comparisons are made between the traditional methods and the proposed one. Finally,

concluding remarks and future work comments are addressed in chapter 6.

13
2. Literature Review

2.1 Basic Concepts

The following are some commonly-used terms and concepts in this thesis:

Feature: Any surface in the machined part (e.g. hole, slot, boss, tab).

Datum: It is a reference feature for machining and measurement.

Dimension: Dimension is the representation of feature size or its location.

Tolerance: The permissible amount of variability in geometry. Limit of size and plus-minus

tolerances are two methods used to specify tolerances. Limit of size means that an upper and

lower limit are given for a specific dimension. As for plus-minus tolerance, a nominal (target

value) followed by a plus-minus expression of a tolerance [Krolikowski (1998)].

Setup: The state of locating and clamping workpiece to be machined.

Fixture unit: A unit that is used to constrain the workpiece from movement during machining.

Size tolerancing (coordinate dimensioning and tolerancing) used to be the only approach for

dimensioning and tolerancing. In this approach, the dimension and its tolerance are

represented by the distance and its variation between two features or points. Although this

approach was found to be successful for many design cases, there were major shortcomings.

These shortcomings showed up because of the increased demand and need for high quality

products. The three main shortcomings are [Krolikowski (1998)]:

1. Coordinate dimensioning does not provide a clear relation between design,

manufacturing and inspection, which could result in different interpretation in

manufacturing and inspecting a part.

2. It does not represent the tolerance zones properly in some cases. An example is that

for a cylindrical feature, the tolerance zone is rectangular.

14
3. The functional requirements (e.g. assembly) in manufacturing the part are not valid.

This is because that the tolerance zone is fixed in size.

In order to remedy these shortcomings when using coordinate tolerancing, long written

comments have to be provided in the design drawings. More practically, Geometric

Dimensioning and Tolerancing (GD&T; ASME Y14.5M-1994) can be used and can solve all

the shortcomings efficiently by:

1. Obtaining clear instructions for inspection and manufacturing (by using the datum

concept).

2. Tolerance zone geometries can be rectangular, circular or cylindrical.

3. Providing clear functional requirements of manufacturing a part by using material

condition modifiers (Maximum Material Condition (MMC), Least Material Condition

(LMC), and Regardless of Feature Size (RFS)).

Geometric tolerances include fourteen types of tolerances that are usually categorized into

five categories; namely: form, orientation, profile, location and runout. Form tolerances

include: flatness, straightness, cylindricity and circularity (roundness). Orientation tolerances

include: parallelism, angularity and perpendicularity. Profile includes: profile of a line and

profile of a surface. Runout tolerances include: circular runout and total runout. Finally,

location tolerances include: position, symmetry and concentricity. They can be further

classified into datum-dependent and datum-independent tolerances. Table 2-1 depicts all the

geometric tolerances, their symbols and their dependencies on datum.

15
Table 2-1: Geometric Tolerances (ASME Y14.5M-1994)

Category Characteristic Symbol Datum Dependency


Flatness
Straightness
Form Never
Cylindricity
Roundness
Parallelism
Orientation Angularity Always
Perpendicularity
Profile of a line
Sometimes
Profile Profile of a surface
Position
Location Symmetry Always
Concentricity
Circular runout
Runout Always
Total runout

Process Capability: The process is considered capable if the process variability is equal or

less than the design specification (tolerance). Usually, it is represented by the Cp index which

is the ratio of design specifications (tolerance, T) to the process variability (6σ).

T USL − LSL
Cp = = (2.1)
6σ 6σ

The USL and LSL are the upper and lower specification limits. Referring to figure 2-1,

considering the design tolerance equals to 6σ (Cp=1) implies that we are satisfied with about

2700 PPM rejects. Nevertheless, Cp index assumes that the process does not drift from the

mean (refer to figure 2-2). Six sigma quality strategy (developed by Motorola in 1980s)

16
assumes ±1.5σ as a typical mean drift. Another index called Cpk considers this drift. It is

defined as:

USL − µ µ − LSL
C pk = min{ , } (2.2)
3σ 3σ

Figure 2-1: With Cp=1, only 2700 part per million (PPM) defects are expected

Figure 2-2: Mean drift in processes

17
2.2 Tolerance Stackup; Definition and Applications

In general, tolerance results from both process tolerance and the tolerance stackup

[Whybrew and Britton (1997)]. The latter is the accumulation (buildup) of error (tolerance) in

a dimension between features resulting from taking operational datums that are different from

the ones indicated in the design specifications. In other words, if the datum indicated in the

design drawings is the one used for locating and clamping, then a stackup-free dimension will

result and there will be no tolerance stackup in this specific dimension. Consequently,

tolerance analysis and tolerance control will not be necessary since the tolerance will depend

solely on the process capability [Huang (1995)]. However, in practice, due to economic

reasons and resource constraints, design datums are not always used as locating and clamping

datums. Therefore, some of the blueprint dimensions will be machined indirectly. Hence, in

most cases tolerance stackup is inevitable.

The way of machining a part determines the stackup in a dimension. There are three main

approaches for machining a part:

1. Chain machining (point-to-point machining): In this approach, the current machined

surface is used as a datum to machine the next surface. This will result in the greatest

accumulation of tolerance.

2. Base-line: This is how parts are machined in a single setup using NC machines. In this

approach, the operational datum is fixed (zeroed) by the coordinate system in the NC

machine for each machining cut. Using this approach decreases the tolerance stackup.

3. Mixed of chain machining and base-line: This happens when parts are machined in

multiple setups.

18
Tolerance allocation is a crucial step in setup planning. Figure 2-3 shows a typical approach

for designing a setup plan. It can be seen that tolerance stackup analysis and tolerance

allocation play important roles in setup planning. Thus, tolerance stackup behavior needs to

be studied carefully and analyzed accurately in order to generate cost-effective setup plans.

During the setup planning, in order to maintain the required tolerances provided in the

blueprints, proper choices of the contributing ones (increasing and decreasing tolerances)

must be made. Achieving this with simulation is possible if we think of the problem in an

opposite way. Rather than providing tolerances for the contributing tolerances to get the

concluding one, the required (concluding) tolerance is provided in order to get tolerances of

contributing ones. Simulation can be run a number of times for a range of the modeled

manufacturing error values to find what the tolerance for each case. A more general

application of this simulation is automating setup planning (tolerance allocation is part of

setup planning). Setup planning can be defined as the act of preparing instructions to machine

a part. Decisions usually taken by the setup planner are: proper datums, machined surfaces,

operations and sequence of operations. The input of the problem is: design requirements and

available resources (tools, machines and fixtures). Essentially, this is an optimization problem

that aims at decreasing: cost and tolerance stackup. Tolerance stackup is part of the cost of

material removal operation. Simulation can be used here to check the goodness of a given

setup plan by examining if the proposed plan leads to acceptable tolerances or not.

19
Setup Plan

- Setup formation
- Datum selection
- Setup sequences

Process
Tolerance
Tolerance
Stackup Analysis
Analysis

No

Tolerance
Allocation

Feasible
Unconstrained Yes Feasible?
Plans

Constrained Optimal Setup


Optimization Plan

Figure 2-3: Typical setup planning approach

2.3 Traditional Analytical Tolerance Stackup Analyses

The general relation of a distance in the x, y and z space can be expressed as following [Lin

and Zhang (2001)]:

d = f ( xi , y j , z k ) (2.3)

Where:

xi: (i=1,…,l) the component dimensions in the X-axis.

yi: (j=1,…,m) the component dimension in the Y-axis.

zk: (k=1,…,n) the component dimension in the Z-axis.

20
Dimension chain (sometimes called tolerance chain) is a closed loop of interrelated

dimensions. It consists of increasing, decreasing links and a single concluding link. In figures

2-4 and 2-5, link i is the increasing link, d is a decreasing link and c is the concluding link.

Apparently, the concluding link c is the one whose tolerance is of interest and which is

produced indirectly. Increasing and decreasing links (both called contributing links) are the

ones that by increasing them, concluding link increases and decreases; respectively.

d c
i

Operational datum
Machined surface

Figure 2-5: Dimension Chain of c, 4 links, 1D


Figure 2-4: Dimension Chain of c, 2 links, 1D

The equation for evaluating the concluding link dimension is [Lin and Zhang (2001)]:

l m
c = ∑i j − ∑ dk (2.4)
j =1 k =1

Where:

∑i: The summation of the increasing link dimensions.

∑d: The summation of the decreasing link dimensions.

j: increasing links index.

k: decreasing links index.

l: number of increasing links.

21
m: number of decreasing links.

For figure 2-4, c can be found as:

c =i−d (2.5)

As for chain in figure 2-5, c can be found as:

c = (i1 + i2 ) − (d1 + d 2 ) (2.6)

2.3.1 Worst-Case Analysis

In worst-case method, the concluding dimension’s tolerance ∆c can be found as following:

l
∂c m
∂c
∆c = ∑ | | ∆i j + ∑ | | ∆d k (2.7)
j =1 ∂i j k =1 ∂d k

Referring to figure 2-5 and equations (2.6 and 2.7), the deviation of the concluding link is:

∆c = ∆i1 + ∆i2 + ∆d1 + ∆d 2 (2.8)

2.3.2 Statistical Analysis

In statistical method, the concluding dimension’s tolerance ∆c can be found as following:

l
∂c m
∂c
∆c = ∑ ( ∂i
j =1
∆i j ) 2 + ∑ (
k =1 ∂d k
∆d k ) 2 (2.9)
j

Here, the tolerance is considered as the difference between two or more independent random

variables (links) which is calculated by adding variances up. Referring to figure 2-5 and

equations (2.5 and 2.9), the deviation of the concluding link c is given by:

∆c = (∆i1 ) 2 + (∆i2 ) 2 + (∆d1 ) 2 + (∆d 2 ) 2 (2.10)

22
2.4 Tolerance Chart

Tolerance chart (developed by Wade in 1983) is a formal, graphical record for a proposed

process plan that is used to verify proposed process plans by identifying tolerance chains and

then checking tolerance stackup in the dimension chains existed in the plan, stock removal

allowances and stock removal tolerance stackup [Whybrew and Britton (1997)]. Firstly,

tolerance chains are identified by the chart and then tolerance stackup are evaluated for the

concluding links. Tolerance chart includes a plenty of chains that are needed to be verified

with design requirements.

Tolerance chart includes the following elements:

(1) Part drawing (at the top of the chart) that includes surfaces of interest,

(2) Operation description (Upper left part),

(3) Working dimension (Upper right part),

(4) Stock removal description (Upper right part),

(5) Chains (Far right),

(6) Machining direction descriptions (Middle of the chart) that are represented by arrows

that start with the operational datum which are represented by donut symbols ( ) and

ends with machined surfaces that are represented by arrow symbol heads ( ),

(7) Blueprint dimensions (Bottom left),

(8) Resultant dimensions (Bottom right) that are needed to be compared with blueprints.

Figure 2-6 shows an example of tolerance chart taken from [Xue and Ji (2002)]. As it can be

noticed in the figure, there are three chains in this plan that must be checked out. After

calculating the tolerance stackup in the critical dimensions, they are compared (in the bottom

23
of the chart) with the design requirements (blueprints). Process plan here satisfies the design

requirements as it can achieve the dimensions and tolerances sought.

Figure 2-6: Example of tolerance chart [Xue and Ji (2002)]

24
3. Simulation-Based Tolerance Stackup Analysis

3.1 Simulation Architecture

Simulation is defined by Kelton (2002) as a board collection of methods and applications to

mimic the real world behavior. We need to tackle the problem of machining tolerance

stackup by simulating the inspection process, after simulating the machining process in terms

of material removal and manufacturing errors. Since manufacturing errors have random

characteristics that can take any probability distribution function (pdf), Monte Carlo

simulation will be the natural choice to solve this problem.

The idea of this simulation is to represent the features of interest by sample points (Figure 3-1

as an example). Then enough number of parts are then virtually machined according to the

intended material removal and the manufacturing errors and inspected according to the

standard CMM (coordinate measuring machine) inspection procedures by tracking the spatial

changes of the features. For more details about the simulation methodology and its

applications, readers should refer to reference [Liu and Huang (2001)]. Simulation is a proper

choice for this problem since other different types of errors can be incorporated in the model.

Furthermore, simulation is not restricted to normal error distributions only; rather, it can take

any probability distribution function (Normal, Uniform, Weibull, Triangular, etc) depending

on the actual error distribution.

25
Figure 3-1: Part representation by sample points

Figure 3-2 is a flowchart that illustrates the general simulation system architecture we are

using in this study. The components of the flowchart are further explained as follows:
NO

Virtual Terminate? Virtual


Setup Plan Sample Plan Error Modeling Stopping criteria
Yes
Machining Inspection

NO

NO Verification?

NO/setup plan YES


enhancement

Validation?

YES

Feasibility?

YES

End

Figure 3-2: System Architecture

26
(1) Setup plan

The flowchart starts with a proposed setup plan for machining the part. Setup plan can be

defined as the instructions for machining a part in order to meet the design requirements by

choosing proper: setup formation, datum and operations sequence. The aim of this planning is

to develop the way of machining a part with the minimum cost and the least tolerance stackup

possible.

(2) Sample plan

In order to represent our parts, we use the same concept used in the coordinate metrology by

representing features by sample points in the space. Since manufacturing processes are far

from perfect, there is no way to yield 100% accurate parts. Therefore, we need to make

representative sample points for the features by choosing proper sample size and sample

point locations. This will be discussed more in details in section 3.3.

(3) Error modeling

Since our simulation is based on simulating manufacturing errors, we need to identify the

contributing error sources that shape up the features in the space. In our model, as it will be

shown later, we adopted the following error sources: (a) cutting tool deviation that includes:

workpiece-tool interaction and cutting tool repeatability and (b) setup error that includes:

fixture unit error and raw part inaccuracies. These errors were categorized (section 3.1.3) and

some evaluation procedures were developed in chapter 5 in case they are not available.

(4) Virtual machining

27
Simulation starts from here by considering a virtual part, shaping its form and orientation in

the space by the sample points and keeping track of the changes of feature representation due

to material removals and manufacturing errors.

(5) Stopping criteria

Validity of the Monte Carlo results depends highly of the number of iterations executed.

Unfortunately, if the number of iterations (number of virtual parts here) is not big enough,

overly misleading results will show up. Therefore, there should be some metrics or criteria

that are used to determine the number of iterations (sample size of the virtual part batch) to

achieve certain accuracy. This will be discussed more in details in section 3.4.

(6) Virtual inspection

After collecting enough data (or sample points/dimensions), tolerances can be evaluated

using the standard methods. This usually includes: datum evaluation, dimensional and

geometric tolerance evaluation. This will be discussed more in details in section 3.2.

(7) Verification

It is the task of ensuring that the simulation is modeled properly. It is also known as

debugging the model. If a bug was found in the code, a review must be done from the start of

the code in the virtual machining part.

(8) Validation

It is the task of ensuring that the simulation model is close enough to the real world behavior.

This is mainly done by conducting real experiment that make physical machining and

inspection for the same part requirements and setup plan and then checking the closeness of

28
the simulation results with the physical ones. As usual, the closeness can be checked by

making statistical inference tests (t and F tests). If a problem was caught at this stage, a

feedback will be given to the manufacturing error model to make another experiment to

check out the manufacturing errors or to lookup at any existed thing in the database.

(9) Feasibility

Simulation checks if the proposed plan is doable using the available resources and taking into

account the constraints.

3.1.1 Monte Carlo Simulation

Monte Carlo methods are numerical methods used to solve probabilistic and deterministic

problems by taking samples from contributing populations and plugging them in the

governing function of the system. Another definition is [Kalos and Whitlock (1986)]: a

numerical stochastic process; that is, it is a sequence of random events.

Monte Carlo Simulation can be further explained as follows: given input random variables

(X1, X2… XN) with their probability distribution functions (pdf’s) and the governing function

that relates them with the output random variable Y=f(X1, X2… XN), approximate behavior of

the output random variable can be found. After enough number of simulation iterations,

distribution of the output random variable can be found (Refer to figure 3-3). Apparently,

increasing number of iterations increases the accuracy of the output. Sample size and point

locations and number of iterations are important parts to be determined when working with

Monte Carlo simulation.

29
Figure 3-3: Monte Carlo Simulation (source: http://www.ymp.gov/documents/ser_b/figures/chap4_2/f04-

174.htm)

3.1.3 Manufacturing Errors

3.1.3.1 Error Categories

Researchers classified manufacturing errors according to different factors (refer to table 3-1).

These factors are:

1. Time. This classification accounts for the error variation with time. Quasi-static errors do

not change considerably (or change slowly) with time such as errors due to dead weights.

Dynamic errors change with time such as cutting tool wear error [Ramesh, Mannan and Poo

(2000)].

2. Randomness. According to this classification, error can be categorized as deterministic

and random errors. Deterministic errors do not have considerable random nature; rather, they

have deterministic dependent output on different independent input parameters such as

30
cutting tool wear. On the other hand, random errors are the ones that change according to

specific probability distribution function (pdf) such as spindle repeatability.

3. Sources of errors. Geometric error sources represent the inaccuracy of surfaces moving

relative to each others. Furthermore, it is believed that it is the biggest contributor in

manufacturing inaccuracy [Ramesh, Mannan and Poo (2000)]. Thermal error accounts for

thermal deformation in the tool because of heat provided by cutting process, machine, people,

thermal memory (from previous environments) and cooling for the coolant. The third major

contributor to inaccuracy of machined part is the cutting-force induced errors that come from

the dynamic stiffness of all components of the machine tool.

4. Errors influence on geometric positions. This classification takes into account the effect

of the error on the finished part. Locating error accounts for the variation between the ideal

datum and the one after locating and clamping. And machining error accounts for the

variation between the ideal position of the machine tool and the actual one [Lin and Zhang

(2001); Huang (1995); Lin, Wang and Zhang (1997)]. This is the classification we adopted in

our model here.

Table 3-1: Manufacturing Error Classification

Factor Categories
- Quasi-static
Time
- Dynamic
- Deterministic
Randomness
- Random
- Geometric
Sources of errors - Thermal
- Cutting force-induced
- Machining (Machine motion
Errors influence on geometric
error)
positions
- Fixture (Setup error)

31
Researchers studied error models either in separate or combined with other types of errors.

Thermal error model and its compensation were studied extensively by [Ramesh, Mannan

and Poo (2000); Okafor, Ertekin and Yalcin (2000); Shuhe, Zhang and Zhang (1997); Yang,

Yuan and Ni (1996); Yang and Lee (1998); Chen, Yuan and Ni (1997); Chen (1996); Wang

et al. (1998); Yang, Yuan and Ni (1996); Elbestawi, Srivastava and Veldhuis (1995);

Krulewich (1998); Chen and Chiou (1995); Ahn and Cao (1999)]. Force-induced error was

studied by [Ramesh, Mannan and Poo (2000), Chiu and Chen (1997)]. Setup error and its

compensation were studied by [Gao, Chase and Magleby (1998); Satyanarayana and Melkote

(2002)]. Different approaches were used to model errors such as: finite element (FE),

artificial neural network (ANN), analytical methods. Table 3-2 shows various manufacturing

error models and their sources and some studies related to each category.

Table 3-2: Manufacturing Error Models

Error Category Author(s) Modeling Approach


Satyanarayana and Melkote 2002 Finite Element Method
Kim et al. 2002 Analytical + on machine measurement
Workholding Error Rong et al. 2001 Vectorial tolerancing
Krishnakumar and Melkote 2000 Genetic algorithm
Hockenberger and De Meter 1996 Quasi static analysis
Ramesh et al. 2002 Artificial neural networks/Bayesian networks
Okafor and Ertekin 2000 Rigid body kinematics
Mize and Ziegert 2000 Artificial neural networks
Yang and Lee 1998 Artificial neural networks + on machine
Machine Tool Error measurement
Wang et al. 1998 Gray system theory
Chen et al. 1997 Rigid body kinematics
Li et al. 1997 Autoregressive
Chen 1996 Artificial neural networks
Hong and Ehmann 1995 Surface shaping system
Okafor and Ertekin 2000 Rigid body kinematics
Cutting Tool Error Schmitz and Ziegert 1999 Vibration frequency analysis
Chen et al. 1997 Rigid body kinematics
Hong and Ehmann 1995 Surface shaping system
Workpiece Error Satyanarayana and Melkote 2002 Finite Element Method
Krishnakumar and Melkote 2000 Genetic algorithm

32
Since the final dimension is either related to one or two features, it is good enough to

consider errors existed in the features of interest. Generally, it is safe to say that

manufacturing inaccuracy can be owed only to two factors; that are: (1) cutting tool deviation

from its theoretical (ideal) path and/or (2) Setup error due to locating, clamping and raw part

inaccuracy. In this thesis, sometimes cutting tool deviation is called machining error (3.1.3.2)

while setup error is sometimes called workpiece error. Setup error can be further divided into

locating/clamping error (3.1.3.3) and raw part error (3.1.3.4). Also, machining error can be

further divided into: cutting tool repeatability and tool-workpiece interaction (refer to figure

3-4).

Dimensional tolerance and most of the geometric tolerances are datum-related. Some of the

geometric tolerances are not datum-related as shown in table 2-1. In the case of datum-

unrelated tolerances (such as flatness, straightness, etc.), cutting tool deviation is enough to

consider whereas in the case of datum-related tolerances (such as dimensional tolerance,

angularity, etc.), cutting tool deviation and setup error must be both considered.

Error Models

Cutting Tool
Setup Error
Deviation
(Workpiece Error)
(Machining Error)

Tool-Workpiece
Repeatability Fixture Unit Error Raw Part Error
Interaction

Figure 3-4: Error models used in this study

33
Figures 3-5 and 3-6 depict setup and machining errors effects on the cutting process.

Referring to figure 3-5, workpiece coordinate system (WCS) deviation from the machine

coordinate system (MCS) causes removing material we do not intend to cut and avoiding

material removal we intend to cut. The inclination in the machined surface shown in figure 3-

6 is caused by the setup error whereas machined surface irregularities represent machining

error effect.

Figure 3-6: Combined effects of setup and


Figure 3-5: Setup Error machining errors

3.1.3.2 Machining Error (Cutting Tool Deviation)

This error accounts for cutting tool path deviation from its idea path. It is assumed in this

study that the deviation is limited to z-coordinate deviation as the cutting tool must travel in

parallel paths. Although, this assumption is valid for prismatic and rotational parts machining,

it is not valid for free-form (sculptured) part machining. This error can be further divided into:

cutting tool repeatability and tool-workpiece interaction error. In this work, we only

considered the tool-workpiece interaction error since the repeatability error is usually

negligible compared to the tool-workpiece interaction error.

34
3.1.3.3 Locating/Clamping Deviation (Fixture Unit Error)

This error accounts for surface in the workpiece deviation from its ideal location due to

clamping and locating. It can be represented by six parameters; namely: translation in x, y

and z and rotation around x, y and z. It is one of the contributors to the setup error.

3.1.3.4 Raw Part Error

Raw part error accounts for part datum inaccuracy contribution to setup error. Part

inaccuracy is represented in our study by the flatness values of the primary, secondary and

tertiary datums. Raw part error and locating/clamping error together establish the setup error.

It can be represented by six parameters; namely: translation in x, y and z and rotation around

x, y and z.

3.1.5 Error Synthesis (Aggregation)

Although, there is a great amount in the literature about machining error modeling and its

compensation, very few researchers attacked the problem of synthesizing the error sources

for multi-operation machining in order to predict the quality of the finished part. This is

because of the complexity of the problem. Yao et al. (2002) developed a desktop virtual-

reality approach to represent the machining and measurement processes by including some

machining error sources in the model. Huang, Zhou and Shi (2002) studied the same problem

35
analytically to determine root-causes of machined part inaccuracy. We argue here about the

need of Monte Carlo simulation use to solve this problem. [Liu and Huang (2001)] presents

the use of Monte Carlo simulation for dimensional accuracy prediction.

3.2 Virtual Inspection

Standardizing and developing accurate methods for evaluation tolerances and datums are

very important. Since different interpretations for the same data can result in different results,

standardizing is so important. Choosing accurate methods for evaluation is important because

if the method is not accurate enough, some good parts can be rejected and some bad parts can

be accepted. It was mentioned previously that tolerance can be categorized into datum-

dependent and datum-independent tolerances. Datum-dependent tolerances evaluation (such

as profile, runout, parallelism, etc) must include datum evaluation. The next two sections

present standard methods to evaluate tolerances when discrete data points for the machined

surface and/or the datum are available.

3.2.1 Datum Evaluation

It is an important task to evaluate the datum in order to find the tolerances related to this

datum. Generally, there are two approaches for evaluating the datum. These approaches are

presented and summarized in [Wilhelm et al. (1996 and1998)] as the following:

36
Least squares (LS) approach: In this method, all the sampled points on a surface (datum)

found by the CMM are fitted and then the fitted plane is translated parallelly to the outmost

point from the material. This method is defined by ISO/WD 5459-3 as the following:

“Location of the datum is defined for planar datums as the plane which is parallel to the least

squares plane and contains the extreme point of the extracted datum feature as measured

from the least squares associated line of the hill in the direction of the outward normal from

the material.”

A procedural interpretation of this definition is as the following [Wilhelm et al. (1996, 1998)]

(refer to figure 3-7):

1. Form the 3D convex hull from the given points.

2. For vertices on the convex hull which are on the surface of the datum feature, not

within the material of the workpiece, fit a least squares plane. These points are the

ones notified by rectangles in figure 3-7.

3. Translate the least squares in the direction of its surface normal away from the

material of the workpiece until the furthest point.

Figure 3-7: Translated least-squares approach

37
Candidate datum set approach: This method is given in ASME Y14.5 standard for GD&T.

According to this approach, two conditions must be met: Firstly, places considered as

candidate datums must be an external set of support. That is, the plane must contact at least

one point on the datum feature while remaining outside of the material part. Secondly, small

facets or sets of support near the center of a part are generally accepted as datums while small

facets near the edges are not [Wilhelm et al. (1996, 1998)].

Procedure given by ASME Y14.5 (refer to figure 3-8):

1. Consider a candidate plane P which is an external set of support for the datum

feature. Let C be the set of contact points between the datum feature and P.

2. Consider an arbitrary line L in P. Orthogonally; project each point in the

feature datum on the line L to get line L’.

3. Divide L’ into three equally spaced segments (unless mentioned).

4. The datum will not be rejected if the contact points C are not located in one

segment.

5. Repeat the same procedures for all the lines in the datum.

38
Figure 3-8: Candidate datum set approach

The ASME Y14.5 standard does not give the details for applying this procedure. Wilhelm et

al. (1996, 1998) proposed methods for evaluating the planar datums and feature of size (FOS).

Wilhelm’s (1998) procedure is as the following:

1. Construct the 3D convex hull for the sampled points. The convex hull consists of facets.

Each facet on the hull that is about the material side of the sampled points is an external set of

support.

2. Each facet is considered as the candidate datum P to be checked. A two dimensional

projection of the candidate facet is taken (refer to figure 3-9).

39
Figure 3-9: 2-D projection of the convex hull [Wilhelm (1998)]

3. Evaluate the facet to be checked by the three region method. Referring to ANSI definition,

the facet corresponds to a plane are the vertices of the facet. All the lines L from the vertices

are constructed. L’ is found as was described before. Figure 3-10 shows an example of a facet

that is not rejected since the contact points between the facet and the feature datum are not

concentrated in one specific region.

Figure 3-10: Non-rejected datum [Wilhelm (1998)]

This approach is mainly a search approach that needs enough number of iterations to find the

appropriate datum. This means that every candidate plane (facet) shall be checked with

infinity number of lines from 0o to 360o. Nevertheless, It is worthy to mention that the lines to

be tested shall not be parallel since parallel lines will lead to the same result; which will

reduce the number of iterations needed.

40
3.2.2 Evaluation Algorithms for GD&T

In order to evaluate a tolerance that depends on a datum, both the datum and the surface must

be evaluated so and the associated tolerance is found accordingly. The allowable variation of

the tolerances in GD&T is based on the envelope principle. The entire surface shall lie

between two ideal envelope features [Zhang (1997)].

CMM data must be further interpreted to evaluate the geometric deviations mathematically.

Usually, this is done by using the least sum of distances fitting, Least Squares fitting (LS) or

the Minimum Zone fitting (MZ). All of them are optimization problems with different

objective functions.

Data fitting in metrology is defined generally by the following equation:

1/ p
1 p
min L p =  ∑ ri  0< p<∞ (3.1)
n i 

Least sum of distances is also known as the median-polish fit. The objective of this function

is equation 1 with p = 1. This fitting is less sensitive to the outliers than the Least Squares

fitting.

Least Squares method is the most widely used approach in CMM data analysis. The objective

function for this approach is given in equation (3.1) with p = 2.

41
The fitted feature for the data points (x, y, z) is called the substitute feature and the geometric

deviation is evaluated as the difference between the maximum and the minimum distances

between the data points and the substitute surface multiplied by 2.

Another widely used method is the minimum zone approach method. The objective function

of this approach (also called two-sided minimax fitting) is given by equation (1) with p → ∞.

The resulting fit is strongly affected by the data outliers. The objective function turns to be as

shown in equation (3.2).

min (max |ei|) for 1 ≤ i ≤ n (3.2)

There is another approach called one-sided minimax fitting which is a constrained

minimization form of the minimum zone. This approach is used to measure the size of the

feature rather than measuring the form deviation. It has two forms, depending if the feature is

internal or external.

For external feature, the optimization problem is:

min (max |ei|) for 1 ≤ i ≤ n (3.3)

Subject to ei ≥ 0

For internal feature, the optimization problem is:

min (max |ei|) for 1 ≤ i ≤ n (3.4)

42
Subject to ei ≤ 0

3.3 Sample Plan

Finding a proper sampling plan for the machined feature to be inspected is a crucial step in

coordinate metrology since the chosen points are considered as the only representative points

of the feature and the other points are overlooked.

In order to have an accurate strategy for sampling points to be measured in a feature, the

minimum sampling size and the best sampling point locations must be found out. In general,

the sampling of a machined feature depends on the machining capability, the part dimensions,

the surface topography, the required tolerance to be found and the accuracy level.

Unfortunately, machined features can never approach the perfect. An awkward solution for

inspecting feature will be by measuring as many points as possible to figure out the shape and

orientation in the space.

Finding the proper sample size (number of points to be inspected) is a major research topic in

the literature. Increasing the number of sample points leads to a more accurate evaluation but

increasing the sample size increases the inspection cost. There are some recommended sizes

for different feature geometries (refer to table 3-3). In table 3-3, the mathematical column

refers to the number of points needed to define the given geometry mathematically and the

recommended values are the ones recommended for measuring the features of the given

geometries. As was mentioned earlier, the more points taken, the more accurate the results

43
are. Another method that can be used to evaluate the sample size is by checking the change of

the results by changing the sample size. Some researchers are proposing using the Shannon-

Nyquist theorem that is used to sample the signals. This theorem states that in order to have a

fair approximation to a wave (in our case is the feature topography), the sampling interval

must be at least double the frequency of the wave.

Table 3-3: Recommended sample size for different geometries [Henzold (1995)]

Feature
geometry Mathematical Recommended

Straight line 2 5
Plane 3 9
Circle 3 7
Sphere 4 9
Cylinder 5 15
Cone 6 15

Concerning the sample point locations, the widely used approaches are: random, uniform

(equidistant), stratified sampling (randomized block or randomized grid), refer to figure 3-11.

In random sampling, the location of each point in the space has the same chance of being

chosen as then others. Uniform sampling distributes the points in the space with fixed

distance between them. Uniform distribution is believed to be very sensitive to periodic

variations in the machined feature. In stratified sampling, the feature is divided into blocks

and a number of sample points are chosen randomly inside each of block. Stratified

distribution has a better coverage of the feature than the random distribution approach. Some

researchers recommend distributions according to low-discrepancy sequences (examples of

these sequences are: Hammersley and Halton-Zaremba sequences) [Woo et al. (1995)]. These

44
sequences were proven mathematically that they cover the area of interest with the least gap

areas possible in the space of interest (refer to figure 3-12).

a) Random b) Uniform (equidistant)

a) Stratified b) Hammersley

Figure 3-11: Sample points located in a square feature using different approaches

Figure 3-12: Sample point locations using random and low-discrepancy methods [Davis and Martin

(1998)]

3.4 Stopping (Terminating) Criteria

It is known that Monte Carlo Simulation is a good solution tool for problems that include

stochastic variables. However, it has the bad reputation as a computationally intensive tool

since it needs a large number of iterations to converge to an acceptable level of accuracy. For

45
our problem here, the number of iterations represents the number of virtually machined part

to be inspected. There are two methods to find a proper amount of iterations (number of

virtually machined and inspected parts); sometimes called terminating criteria. Making

approximate statistical calculations to find the sample size is the first method. The second one

is by using empirical methods by considering a tolerance band.

Statistically, the minimum number of iterations can be calculated as follows. Suppose that

when running the simulation for no iterations, the half width (ho) of the confidence interval is

given by the following equation when sample standard deviation (so) is known:

so
ho = t n −1,1−α / 2
no (3.5)

When we want to achieve half confidence interval (h), then the number of termination

iterations can be calculated using the following equation:

2
so
n = t 2 n −1,1−α / 2 2
h (3.6)

However, there is an apparent difficulty that the right hand side of the equation depends on a

prior knowledge of n. In order to overcome this problem, we can replace the t random

variable with standard normal critical values as shown in the following equation (this is valid

when the sample size is over 30).

s2
n ≅ z 21−α / 2
h2 (3.7)

46
An easier but different approximation is given by the following equation [Kelton, Sadowski

and Sadowski (2002)]:

2
h
n ≅ no o2
h (3.8)

Cvetko, Chase and Magleby (1998) developed new metrics to evaluate their simulation. One

method presented in their paper is by benchmarking the results of the simulation for a big

sample size (e.g. 1 billion). The objective of benchmarking the results at such a big number is

to evaluate the performance of the simulation at different sample sizes. When there are no

change in the fist four moments (mean, variance, skewness and kurtosis), the sample size of

the number of iterations is chosen (refer to figure 3-13). From the figure, it can be seen that at

1 million iterations, the results are roughly accurate by 95%.

Figure 3-13: Benchmarked results at 1 billion iterations [Cvetko, Chase and Magleby (1998)]

47
4. Manufacturing Error Evaluation

The objective of this chapter is to introduce the requirements and procedures of experiments

and algorithms for evaluating manufacturing error; which are the inputs of the simulation.

This will be required in case the manufacturing errors are not available. Afterwards, the error

distributions and their parameters (normal, uniform …) can be plugged in the simulation to

get the results. The discussed manufacturing errors here are machining and setup errors. As it

was previously mentioned, machining error can be further classified as: cutting tool

repeatability and cutting tool-workpiece interaction error. And setup error can be further

classified as: fixture unit and workpiece irregularities errors.

4.1 Machining Error Evaluation

If the machining error (cutting tool deviation) is not available, an experiment must be

conducted to evaluate it. The following are the requirements and the procedure of a proposed

experiment we developed to evaluate machining error for a specific CNC machine. This

experiment can be used to evaluate both the: cutting tool repeatability and cutting tool-

workpiece interaction error.

Requirements:

1. A prismatic Aluminum part.

2. A fixture unit.

3. CNC milling machine.

4. Magnetic dial indicator.

48
Procedure:

1. Clamp the part properly.

2. Mill the surface with 2 mm depth material-removal (however, depth of cut can be

chosen by the experimenter).

3. Measure enough number (say 30) of point heights (z-coordinate) using dial indicator

without removing the part from the fixture and without removing the fixture unit from

the machine table (refer to figure 4-1). Fill the measurements of the dial indicator and

the nominal heights of the part after machining in table 4-1. We need not to remove

the part from the fixture unit to isolate the effect of the fixture unit error here and take

on-machine measurements (OMM). Notice that the dial indicator (figure 4-1) is

attached to the machine body by a magnet and does not move. Machine table is

moved in order to make the measurements.

4. Repeat from step 2 enough number of times (say 30).

Table 4-1: Data Collection

Trial/dial
Nominal
indicator 1 2 … 30
Height
measurement
1
2

30

5. Fill the table 4-2 with the differences between the measured heights and the nominal

heights from table 4-1.

49
Cutting Tool

Dial Indicator

Prismatic
Workpiece

Fixture Unit

Machine Table

Figure 4-1: Dial Indicator measuring machined surface

Table 4-2: Machining Error Data

Trial/dial
indicator
1 2 … 30
measurement-
nominal height
1
2

30

6. Fit the data in table 4-2using a proper distribution by finding the first four moments (mean,

variance, skewness and kurtosis.)

50
Notes:

(1) Distributions of the measurements for the same machined surface (horizontal data points

in table 4-2) represent error due to cutting tool-workpiece interaction. On the other hand,

distributions of the height deviation measurement after each machining (vertical data points

in table 4-2) represent cutting tool repeatability.

(2) It will be much better to use on-machine-measurement (OMM) in order to take

measurements here rather than using a magnetic dial indicator. The OMM is a measuring

system that can be used as coordinate measuring machines by replacing the cutting tool with

a probe. This will increase the accuracy here as we will no longer depend on the magnet to

sustain the dial indicator from movement.

4.2 Locating/Clamping Error (Fixture Unit Error) Evaluation

If the locating/clamping error is not available, an experiment must be conducted to evaluate it.

The following are the requirements and the procedure of a proposed experiment we

developed to evaluate this type of error for a specific fixture unit shown in figure 4-2.

51
Fixture
Unit

Prismatic
Workpiec

Figure 4-2: Fixture Unit with the workpiece

Requirements:

1. A prismatic standard smooth part. The part must be accurate enough (say at most 2

microns flatness values for all the surfaces.)

2. A fixture unit.

3. Coordinate Measuring Machine (CMM).

Procedure:

1. Choose three adjacent faces of the prismatic part (figure 4-3).

2. Mark the upper corner of the edge formed by three surfaces as workpiece origin.

52
Figure 4-3: Part Surfaces

3. Clamp the part in the fixture. Locate the fixture unit itself on the CMM table in order

to take measurements.

4. Establish workpiece coordinate system by measuring enough number of points (say

10 random points) for each of the following surfaces: xy, yz and zx.

5. Unclamp the part, remove it from the fixture and clamp it back in the same position.

6. Repeat step 4.

7. Deviation between coordinate system established in step 4 and 6 will give the

translational and rotational errors due to fixture errors.

8. Repeat from 5 enough number of times (say 30 times) using the same part and enter

data in table 4-3.

9. Fit the data in table 4-3 using a proper distribution by finding the first four moments

(mean, variance, skewness and kurtosis) for the six parameters.

53
Table 4-3: Fixture unit error parameters

tx ty tz α β γ

30

4.3 Raw Part Error Evaluation Algorithm

Flatness values of primary, secondary and tertiary datums are the input of the raw part

inaccuracy to the setup error in our model. The algorithm used to evaluate flatness values

contribution to the setup error is:

1. Generate three points to represent each datum using the given flatness values. The

distribution to generate the values from is normal with mean = 0 and standard

deviation = Flatness/6. The points must be 1/3 or more the surface dimension apart

from each others in each surface.

2. Transform the generated sample points into machine coordinate system using the

found values for fixture unit deviations (clamping/locating error for a smooth part).

3. Find LSE planes for each datum.

4. Find the point of intersection between the three fitted planes. Consider the intersection

point as the translational error parameters.

5. Find the normal vectors of each datum (n1, n2, n3) using the following equations:

θ x = cos −1 (n1 .[0,0,1])


θ y = cos −1 (n2 .[0,1,0]) (4.1)
θ z = cos −1 (n3 .[1,0,0])

54
6. Repeat from step 1 for enough number of times (here we recommend at least 100,000

times). It is worthy to mention that the condition in step 1 will reduce the number of

iterations to approximately 6%.

7. Evaluate means and variances for each angle deviation.

In order to evaluate the performance of the proposed algorithm, an experiment was conducted

in Delphi Automotive System. The output of the program must match the experimental

results in order to consider the proposed algorithm representative. The input of the program

will be the flatness values of the primary, secondary and tertiary datums. The following are

the requirements and the procedure of the experiment.

Requirements:

1. Coordinate measuring machine (CMM).

2. A fixture unit.

3. Smooth part that is accurate enough (say average flatness less than 2 microns). It is

assumed that there is no effect of the smooth part flatness on the setup error.

4. Rough part.

Figures 4-4 and 4-5 show CMM, fixture unit and the part.

Procedure:

1. Evaluate the fixture unit error in terms of the 3 translational and 3 rotational

deviations using a smooth part (stone part) using exactly the same procedure

mentioned in section 4.2.

55
Probe
Fixture
Unit

Workpiec

Figure 4-4: Coordinate measuring Figure 4-5: Fixture unit with CMM probe
machine (CMM)

2. Evaluate primary, secondary and tertiary datum flatness values for the rough part

using CMM (input of the program).

3. Find the rotational and translational deviations for the rough part using the same

procedure described in section 4.2.

The output of the program (simulation) was found to be consistent and close to the

experimental results. However, although the results found look fairly close and stable,

sometimes results show some very different behavior. In other words, sometimes results are

truly misleading. This can be justified by insufficient iterations of the simulation. A sample of

the program output and the experimental results are shown in tables 4-4 and 4-5. It is clear

that the simulation output and the experimental results are statistically the same since the p-

value is very high for all the cases except for θx in table 4-5.

Table 4-4 data are for extremely smooth part (all flatness values are 0). And table 4-5 is for a

rough part. Therefore, the following null hypothesis cannot be rejected:

56
H 0 : σ 2 simulation = σ 2 exp eriment
(4.2)
H 1 : σ 2 simulation ≠ σ 2 exp eriment

Table 4-4: Variance comparison between simulation and experiment for smooth part

Rotational Experimental Simulation Results Statistic p-value


parameters Results m = 6117 (df=6116) F
n = 30 (df=29)
θx 6.761e-006 4.85448499311699e-006 1.39273 0.1569
θy 1.925e-006 2.15862432856938e-006 0.891772 0.5963
θz 3.945e-006 2.91489690848275e-006 1.353393 0.19544

Table 4-5: Variance comparison between simulation and experiment for a rough part

Rotational Experimental Simulation Results Statistic p-value


parameters Results m = 6207 (df=6206) F
n = 30 (df=29)
θx 0.001119036304 0.00260209785491036 0.430051584 0.00014
θy 0.000644855236 0.000613310910554736 1.051432846 0.78038
θz 0.003803312241 0.00561939444032142 0.676818878 0.09514

57
5. Illustrative Examples

The following are three illustrative examples that are used to evaluate tolerance stackup using

our simulation. The part in the first example is virtually machined within single setup since

there is no datum change during machining. In the second example, three setups for the same

part in example 1 are considered. Two setups are considered for the third example.

5.1. Example 1: Two Machining Operations (Within One Setup)

Figure 5-1 shows the design drawing of an example part with its dimensions and tolerances.

The sequence of operations and the working dimensions of machining are shown in the left of

the figure. The simulation is conducted to find out the stackup behavior of the dimension

between surfaces f2 and f4 after machining them in a single setup using the same datum, f0.

Figure 5-1: Example 1 (design requirements and the machining sequence)

The errors included in the simulation are (e.g. provided from manufacturing capability

database):

58
• Machining error = N (µ=0, σ2=0.038 m2)

• Rotational setup error =U (-0.003 m, 0.005 m)

• Translational setup error =U (-0.004o, 0.010o)

Figure 5-2 depicts the output of the simulation conducted for 500 virtual parts (500 iterations).

Notice that the variation of the concluding link was found to be less than the other links,

which contradicts to what were expected using traditional methods. According to traditional

methods of evaluating the tolerance stackup, tolerance of the concluding link (distance

between features f2 and f4) should be the summation of the two deviations of the other links in

the chain in the worst-case scenario and the square root of the sum of squares of the two

deviations in statistical analysis. Actually, getting such a lower variation in the concluding

link is justified in our point view since these two features are machined in the same setup.

There is no tolerance stackup in this case, as has been demonstrated in [Huang (1995)].

Machining error is the only error that causes the variation here. There is no contribution from

the setup error. Table 5-1 shows a comparison of the tolerance stackup evaluation using the

three methods.

59
Figure 5-2: Example 1 results

Table 5-1: Tolerance stackup evaluation comparison for example 1

δ02 0.057648
δ04 0.052874
Worst Case Statistical Simulation
δ24 0.1105 0.0782 0.0126

5.2. Example 2: Four Machining Operations (In Three Setups)

The example is an extension of the first one (refer to figure 5-3). It involves four machining

operations with two changes of the machining datums, a total of three setups. Changing

datums will result in a tolerance chain. The errors included in the simulation of example 2 are

the same as that in example 1.

Figure 5-4 depicts the output of the simulation for 500 virtually machined parts (iterations).

Again, the results here do not agree or even close to either the worst-case or the statistical

method. Table 5-2 summarizes the results.

60
Figure 5-4: Example 2 output
Figure 5-3: Example 2 (design requirements
(The simulation output for the

distributions of the distances

between surfaces)

Table 5-2: Tolerance analysis results for example 2

Worst-Case Statistical Simulation


δ24 0.1105 0.0780 0.0528

5.3. Example 3: ABS part

The third example (shown in figure 5-5) is a housing part used by Bosch. A simplified

drawing of the final product dimensional requirements is shown in figure 5-6.

61
Figure 5-5: Example 3; ABS (Antiblock System) housing, Bosch (Source: http://www.wzl.rwth-
aachen.de/WM/SIMON/deliverables/DA0/DA0_02D.htm)

Figure 5-6: ABS dimensional requirements

A simulation was conducted according to the setup plan described in figure 5-7. The plan

includes two setups. The part includes six surfaces of interest that are numbered from 1 to 6.

Milling is the process used to machine the surfaces. Hole drilling is not included in the setup

plan and simulation since it does not have an effect on the tolerance chain of concern. Figure

5-8 shows a tolerance chart of the part in order to predict tolerance stackup in the tolerance

chains. Here, we are interested in the dimension shown in line 8 in the tolerance chart as a

concluding link. The contributing links are shown in lines: 7, 4 and 1.

62
As was mentioned earlier, determining the number of iterations for the simulation is a crucial

task in Monte Carlo simulation. Here, we benchmarked the results of the simulation at

100,000 to calculate the errors in the first two moments when having less sample size.

Actually, 10,000 iterations are considered large enough by most of Monte Carlo parishioners

[Cvetko, Chase and Magleby (1998)]. The first two moments (mean and variance) at 100,000

iterations are shown in table 5-4 for dimensions in lines: 1, 4, 7 and 8.

Table 5-3: Simulation results at 100,000 iterations

Mean Variance
L1 84.9983246659704 0.000113812363943454
L4 59.9943255754447 0.000114461403303197
L7 49.9926265647646 0.000118204861229129
L8 74.9965883885753 0.000118607406206314

The following were the inputs to the simulation:

Flatness of the raw part: 0.05 mm (Flatness is considered to be representative for the raw part

error.)

Machining error (Cutting Tool deviation) ~ N (0, 0.00752)

Rotational setup error ~ U (-0.002, 0.005) degrees

Translational setup error ~ U (-0.0015, 0.005) mm

63
4,5,6
4 4
4,5,6
5 5
6

100
z z z z

y 1,2,3 y 1,2,3 y 1,2,3 y 1,2,3


x x x x

90
160

RAW PART 1 RAW PART, MATERIAL REMOVAL: 15 2 RAW PART, MATERIAL REMOVAL: 15 3 RAW PART, MATERIAL REMOVAL: 10

1,2,3
1 1 1
2 2 2
3 3

z z z z

y 6 y 6 y 6 y 6

x 5 4 x 5 4 x 5 4 x 5 4

4 RAW PART, MATERIAL REMOVAL: 10 5 RAW PART, MATERIAL REMOVAL: 15 6 RAW PART, MATERIAL REMOVAL: 10 FINAL PART

Figure 5-7: ABS part setup plan

A comparison between the results of simulation, worst-case and statistical methods in finding

concluding link tolerance stackup is shown in table 5-4. Again, Monte Carlo simulation

results were found to be less than both the traditional methods’. The ratio of simulation

tolerance stackup to the worst case tolerance stackup was found to be 0.34 and the ratio of

simulation tolerance stackup to the statistical method tolerance stackup was found to be 0.59.

Table 5-4: Tolerance evaluation using the three approaches

Standard
Tolerance=6σ
Deviation
L1 0.010668288 0.064009727
L4 0.010698664 0.064191982
L7 0.010872206 0.065233235
L8 0.010890703 0.065344216

Worst Case 0.193434944


Statistical 0.111683618
Simulation 0.065344216

64
Figure 5-8: Tolerance Chart of ABS part

Figures 5-9 shows the dimension histograms of the concluding link (L8) and the contributing

links (L1, L4 and L7). Number of iterations required to achieve certain accuracy can be

predicted from figure 5-10. This figure shows means and standard deviations values predicted

using the simulation in terms of number of iterations (x axis). Clearly, 4000 iterations seem to

have very close results to the 100,000 iterations. Therefore, 4000 iterations can be considered

as proper choice for the sample size virtually machined parts (iterations.)

65
Figure 5-9: Dimensions histogram using simulation

Figure 5-10: Progress of results with sample size increase

Suppose that we need to maintain 0.066 mm for dimension shown in line 8 in the tolerance

chart (figure 5-8). Then, we will need to allocate proper tolerances for the concluding links

(dimensions shown in lines: 1, 4 and 7 in the same figure). According to our simulation,

assigning 0.060 mm for each contributing link will be good enough to meet what we need.

However, if we need to make the allocation using the worst case and statistical methods,

0.022 mm and 0.038 mm will be needed for each contributing link.

66
If we choose processes that are capable to achieve our simulation requirements, then we will

be satisfied with 2,700 parts per million (PPM) rejects when having the process capability

index equals to 1. However, if we use the same process considering the traditional methods,

much more rejects per million will be expected (refer to table 5-5 and figure 5-11). This

shows the importance of having less conservative method for tolerance allocation.

Figure 5-11: Rejection areas comparison when allocating concluding links tolerance using worst case,
statistical and simulation methods

Table 5-5: Part per million (PPM) rejections comparison when allocating tolerance using worst case,

statistical and simulation methods

Worst Case Statistical Simulation


Tolerance 2.02σ 3.5σ 6σ
Cp 0.337 0.583 1
PPM rejects 307,728 76,727 2,700

67
6. Concluding Remarks and Recommendations

6.1 Summary

1. A new method for evaluating tolerance stackup using Monte Carlo simulation was

developed in this work. Cutting tool deviation and setup error were simulated along with

machining and inspecting processes for virtual parts. This method gives less conservative

results compared to the traditional ones (worst case and statistical methods) and overcomes

the drawbacks discussed earlier.

2. Generally, it is safe to say that manufacturing inaccuracy can be owed only to two factors;

that are: (1) cutting tool deviation from its theoretical (ideal) path and/or (2) deviation of the

workpiece itself due to locating, clamping and raw part inaccuracy.

3. The only practical way for checking the possibility of getting the design requirement out of

proposed process plan is by simulating the manufacturing and inspection processes.

4. Tolerance analysis in machining should be carried out by simulating the manufacturing

errors involved, which are the root causes for dimensional and geometric inaccuracy of

machined components.

5. Worst-case and statistical methods give more conservative results compared to the

proposed one. Overestimating tolerance stackup could result in rejecting good process plans

that should be accepted. Therefore, accurate evaluation of the stackup can lead to cost-

effective process plans.

68
6. Simulation is a proper choice for this problem because of its complexity. Furthermore,

simulation is not restricted to normal error distributions only; rather, it can take any

probability distribution function (Normal, Uniform, Weibull, Triangular, etc) depending on

the actual error distribution. It is precious to mention that even though statistical tolerances

are assumed to be normally distributed; a lot of evidences in the real world defy this

assumption [Lin, Wang and Zhang (1997)].

7. Monte Carlo simulation is believed to be a powerful tool to solve problems that include

stochastic variables. However, the main critique to this method is the need for a quite large

number of iterations to converge to accurate enough results. 10,000 iterations are considered

as large enough by most Monte Carlo practitioners [Cvetko, Chase and Magleby (1998)]. In

our work here, we benchmark the results at large iterations size (like 100,000 or 1 million)

and consider these results as absolutely accurate ones. Afterwards, we calculate the errors by

increasing the sample size. We can choose a sample size that has close results to the

benchmarked ones.

6.2 Recommendations for Future Works

1. Simulation validation and verification increase user’s confidence of the results’ accuracy.

Verification can be defined as the assessment of how close simulation results are to the

conceptual model. In other words, it is the task of ensuring that the simulation was built

accurately as the modeled indeed wanted.

69
Validation is used to ensure that the simulation model matches accurately enough with the

real world behavior. In our problem here, some experiments are recommended to be

conducted for validation. Some experiments must be done to determine simulation input in

case they are not already known. These inputs are the considered manufacturing errors

included in the simulation. Afterwards, a real machining must be done for a large enough

number of parts according to the same setup plan adopted in the simulation to evaluate the

tolerances of the dimensions. The output of the simulation experiment will match the results

of the experiment if the simulation model is valid. Typically, statistical inference tests can be

good to test the closeness between simulation results and the real world behavior.

2. Monte Carlo simulation is known as a computationally extensive tool of calculation.

Therefore, developing more efficient methods in terms of calculation time could be a

valuable future work.

70
Bibliography

[1] Ahn, K. G. and Cho, D. W., “In-Process Modeling and Estimation of Thermally

Induced Errors of a Machine Tool During Cutting,” International Journal of Advanced

Manufacturing Technology, Vol. 15, Issue: 4, April 27, 1999, pp. 299 - 304.

[2] ANSI Y14.5 M-1994, Mathematical Definition of Dimensioning and Tolerancing

Principles.

[3] Banks, J., Discrete-Event System Simulation, Prentice Hall, Third Edition, 2001.

[4] Carr, D., A Comprehensive Method for Specifying Tolerance Requirements for

Assemblies, Master’s thesis, Brigham Young University, 1993.

[5] Chase, K. W. and Parkinson, A. R., “A Survey of Research in the Application of

Tolerance Analysis to the Design of Mechanical Assemblies,” Research in

Engineering Design, Vol. 3, 1991, pp.23-37.

[6] Chase, K.; Gao, J. and Magleby, S., “General 2-D Tolerance Analysis of Mechanical

Assemblies with Small Kinematic Adjustments,” Publications of the Association for

the Development of Computer-Aided Tolerancing Systems (ADCATS), Brigham

Young University, 1995.

[7] Chen, J. S. and Chiou, G., “Quick testing and modeling of thermally-induced errors of

CNC machine tools,” International Journal of Machine Tools and Manufacture, Vol.

35, Issue: 7, July, 1995, pp. 1063-1074.

[8] Chen, J. S.; Yuan, J.; Ni, J., “Thermal error modeling for real-time error

compensation,” Precision Engineering Vol. 20, Issue: 2, April, 1997, pp. 151.

[9] Chen, J.-S. , “Fast calibration and modeling of thermally-induced machine tool errors

in real machining,” International Journal of Machine Tools and Manufacture, Vol. 37,

Issue: 2, February, 1997, pp. 159-169.

71
[10] Chen, J.-S., “A study of thermally induced machine tool errors in real cutting

conditions,” International Journal of Machine Tools and Manufacture, Vol. 36, Issue:

12, December, 1996, pp. 1401-1411.

[11] Chen, J.-S., “Computer-aided accuracy enhancement for multi-axis CNC machine

tool,” International Journal of Machine Tools and Manufacture, Vol. 35, Issue: 4,

April, 1995, pp. 593-605.

[12] Chen, X.B.; Geddam, A. and Yuan, Z.J., “Accuracy Improvement of Three-Axis CNC

Machining Centers by Quasi-Static Error Compensation,” Journal of Manufacturing

Systems, Vol. 16, Issue: 5, 1997, pp. 323-336.

[13] Chiu, W.M. and Chan, K.W., “Design and testing of piezoelectric actuator-controlled

boring bar for active compensation of cutting force induced errors,” International

Journal of Production Economics, Vol. 51, Issue: 1-2, August 15, 1997. pp. 135-148.

[14] Craig, C., “Dimensional Management versus Tolerance Assignment,” Assembly

Automation, Vol. 16, No. 2, 1996, pp. 12-16.

[15] Cvetko, R.; Chase, K. W. and Magleby, S. P., “New Metrics for Evaluating Monte

Carlo Tolerance Analysis of Assemblies,” Proceedings of the ASME International

Mechanical Engineering Conference and Exposition, Anaheim, CA, Nov. 15-20, 1998.

[16] Davis, T. and Martin, R., “Low-Discrepancy Sequences for Volume Properties in

Solid Modeling,” CSG98, 1998, pp. 139-154.

[17] Drysdale, R.; Jerard, R. and Schaudt, B., “Discrete Simulation of NC programming,”

Algorithmica, Vol. 4, 1989, pp. 33-60.

[18] Elbestawti, M. A.; Srivastava, A. K. and Veldhuis, S. C., “Modeling geometric and

thermal errors in a five-axis CNC machine tool,” International Journal of Machine

Tools and Manufacture, Vol. 35, Issue: 9, September, 1995, pp. 1321-1337.

[19] Feng, S. and Hopp, T., “A Review of Current Geometric Tolerancing Theories and

72
Inspection Data Algorithms,” NIST, 1991.

[20] Gao, J. ; Chase, K. and Magleby, S., “Generalized 3-D tolerance analysis of

mechanical assemblies with small kinematic adjustments,” IIE Transactions, Vol. 30,

Issue: 4, 1998, pp. 367-377.

[21] Gao, J., Chase, K. W. and Magleby, S. P., “Comparison of Assembly Tolerance

Analysis by the Direct Linearization and Modified Monte Carlo Simulation Methods,”

Proceedings of the ASME Design Engineering Technical Conferences, Boston, MA,

1995, pp. 353-360.

[22] Hockenberger, M. J. and De Meter, E. C., “The Application of Meta Functions to the

Quasi-Static Analysis of Workpiece Displacement within a Machining Fixture,”

Journal of Manufacturing Science and Engineering, Vol. 118, 1996, pp. 325-331.

[23] Hong, M. and Ehmann, K., “Generation of engineered surfaces by the surface-shaping

system,” International Journal of Machine Tools and Manufacture, Vol. 35, Issue: 9,

September, 1995, pp. 1269-1290.

[24] Huang, Q.; Zhou, S. and Shi. J., “Diagnosis of Multi-Operational Machining Processes

by Using Virtual Machining,” Robotics and Computer-Integrated Manufacturing, Vol.

18, Isuues: 3-4, 2002, pp. 233-239.

[25] Huang, S. H. and Zhang, H.-C., “Use of Tolerance Chart for NC Machining,” Journal

of Engineering Design and Automation, Vol. 2, No. 1, 1996, pp.91-104.

[26] Huang, S., A Graph-Matrix Approach to Setup Planning in Computer-Aided Process

Planning (CAPP), PhD dissertation, Department of Industrial Engineering Texas Tech

University, 1995.

[27] Kalos and Whitlock, Monte Carlo Methods, Vol. 1, Wiley Sons, 1986.

[28] Kelton, W. D.; Sadowski, R. P. and Sadowski, D. A., Simulation with Arena,

McGraw-Hill, second edition, 2002.

73
[29] Kim, S. H.; Ko, T. J. and Ahn, J. H., “Elimination of Settling Error Due to Clamping

Forces in On-Machine Measurement,” International Journal of Advanced

Manufacturing Technology, Vol. 19, 2002, pp:573-578.

[30] Kim, W. and Ramanam, S., “On the selection of flatness measurement points in

coordinate measuring machine inspection,” International Journal of Machine Tools

Manufacture, Vol. 40, 2000, pp. 427-443.

[31] Krishnakumar, K. and Melkote, S., “Machining fixture layout optimization using the

genetic algorithm,” International Journal of Machine Tools and Manufacture, Vol. 40,

Issue: 4, March, 2000, pp. 579-598.

[32] Krolikowski, A., Fundamentals of Geometric Dimensioning and Tolerancing, Second

Edition, Delmar Publisher, 1998.

[33] Krulewich, D. A., “Temperature integration model and measurement point selection

for thermally induced machine tool errors,” Mechatronics, Vol. 8, Issue: 4, June, 1998,

pp. 395-412.

[34] Landry, M.; Malouin, J.L. and Oral, M., “Model Validation in Operations Research.”

European Journal of Operational Research, Vol. 14, No. 3, 1983, pp 207-220.

[35] Law, L., Multivariate Statistical Analysis of Assembly Tolerance Specifications,

Master’s thesis, Brigham Young University, 1995.

[36] Lee, G.; Mou, J. and Shen, J., “Sampling Strategy Design for Dimensional

Measurement of Geometric Features Using Coordinate Measuring Machine,”

International Journal of Machine Tools Manufacture, Vol. 37, No. 7, 1997, pp. 917-

934.

[37] Li, S. ; Zhang, Y. and Zhang, G., “A study of pre-compensation for thermal errors of

NC machine tools,” International Journal of Machine Tools and Manufacture, Vol. 37,

Issue: 12, December, 1997, pp. 1715-1719.

74
[38] Lin E. and Zhang, H.-C., “Theoretical Tolerance Stackup Analysis Based on

Tolerance Zone Analysis,” The International Journal of Advanced Manufacturing

Technology, Vol. 17, 2001, pp. 257-262.

[39] Lin, S.; Wang, H.-P. and Zhang, C., “Statistical Tolerance Analysis Based on Beta

Distribution,” Journal of Manufacturing Systems, Vol. 16, No. 2, 1997, pp. 150-158.

[40] Liu, Q. and Huang, S. H., “Prediction of Component Dimensional and Geometric

Accuracy through Manufacturing Error Analysis,” Transactions of the North

American Manufacturing Research Institution, SME, Vol. 29, 2001, pp. 525-532.

[41] Liu, S. C. and Hu, S. J., “An Offset Finite Element Model and its Applications in

Predicting Sheet Metal Assembly,” International Journal of Machine Tools

Manufacturing, Vol. 35, No. 11, 1995, pp. 1545-1557.

[42] Liu, Z.-Q., “Repetitive Measurement and Compensation to Improve Workpiece

Machining Accuracy,” International Journal of Advanced Manufacturing Technology,

Vol. 15, Issue: 2, February 24, 1999, pp. 85 - 89.

[43] Mize, C. and Ziegert, J., “Neural network thermal error compensation of a machining

center,” Precision Engineering, Vol. 24, Issue: 4, 2000, pp. 338-346.

[44] Musa, R. A. and Huang, S. H., “3D Tolerance Stackup Analysis,” Proceedings of the

Industrial Engineering Research Conference (Portland, Oregon), May, 2003.

[45] Okafor, A.C. and Ertekin, Y. M., “Derivation of machine tool error models and error

compensation procedure for three axes vertical machining center using rigid body

kinematics,” International Journal of Machine Tools and Manufacture, Vol. 40, Issue:

8, June, 2000, pp. 1199-1213.

[46] Ramesh, R.; Mannan, M. A. and Poo, A. N., “AI-Based Classification Methodologies

for Modeling Machine Tool Thermal Error,” Transactions of the North American

Manufacturing Research Institution/SME, 2002, pp. 239-246.

75
[47] Ramesh, R; Mannan, M.A and Poo, A. N., “Error Compensation in Machine Tools –

A review, Part I: Geometric, Cutting-Force Induced and Fixture Dependent Errors,”

International Journal of machine tools and manufacture, Vol. 40, 2000, pp. 1235-1256.

[48] Ramesh, R; Mannan, M.A and Poo, A. N., “Error Compensation in Machine Tools –

A review, Part II: Thermal Errors,” International Journal of machine tools and

manufacture, Col. 40, 2000, pp. 1257-1284.

[49] Rong, Y.; Hu, W.; Kang, Y.; Zhang, Y. and Yen, D., “Locating Error Analysis and

Tolerance Assignment for Computer-Aided Fixture Design,” International Journal of

Production Research, Vol. 39, No. 15, 2001, pp. 3529-3545.

[50] Satyanarayana, S. and Melkote, S., “Determination of Clamping Force Based on

Minimization of Workpiece Elastic Deformation,” NAMRC 2002, Society of

Manufacturing Engineers (SME).

[51] Schmitz, T. and Ziegert, J., “Examination of Surface Location Error Due to Phasing of

Cutter Vibration,” Precision Engineering, Vol. 23, 1999, pp. 51-62.

[52] Thimm, G.; Britton, G. and Cheong, F., “Controlling Tolerance Stacks for Efficient

Manufacturing,” The International Journal of Advanced Manufacturing Technology,

Vol. 18, 2001, pp. 44-48.

[53] Wade, O. R., Tolerance Control, Chapter 2 in Tool and Manufacturing Engineers

Handbook, Vol. 1, Machining, Drozda, T. J and Wicks, C. (editors), Society of

Manufacturing Engineers (SME), Dearborn, MI, 1983.

[54] Wang, Y.; Zhang, G.; Moon, K. and Sutherland, J. W., “Compensation for the thermal

error of a multi-axis machining center,” Journal of Materials Processing Technology,

Vol. 75, Issue: 1-3, March 1, 1998, pp. 45-53.

76
[55] Wang, Z., Modeling and Sampling of Work piece Profiles for Form Error Evaluation,

M.S thesis, University of Cincinnati, 2000.

[56] Whybrew, K. and Britton, G. A., “Tolerance Analysis in Manufacturing and Tolerance

Charting,” Advanced Tolerancing Techniques (edited by H.-C. Zhang), Chapter 2,

John Wiley & sons, INC., 1997.

[57] Wilhelm, R.G. and Chui, B., “Computational Metrology for the Set of Candidate

Datum Reference Frames,” Transactions of the North American Manufacturing

Research Institute of SME, Vol. 24, 1996, pp.181-186.

[58] Wilhelm, R.G.; Bapat, S.; Reddy, P.V.R. and Yau, H-T., “Computational Metrology

of Datums: Algorithms and a Comparative Study,” Proceedings of the 7th Annual

Industrial Engineering Research Conference, Alberta, Canada, May 9-10, 1998.

[59] Wilhelm, R.G., “Mathematical Definition of Geometrical Tolerances,” Technical

Papers of the Society of Manufacturing Engineers (SME), 1997.

[60] Woo, T.C.; Liang, R.; Hsiang, C.C. and Lee, N. K., “Efficient Sampling for Surface

Measurement,” Journal of Manufacturing Systems, Vol. 14, No. 5, 1995, pp. 345-354.

[61] Xue, J. and Ji, P., “Identifying tolerance chains with a surface-chain model in

tolerance charting,” Journal of Materials Processing Technology, Vol. 123, issue: 1,

2002, pp. 93-99.

[62] Yang, M. and Lee, J., “Measurement and prediction of thermal errors of a CNC

machining center using two spherical balls,” Journal of Materials Processing

Technology, Vol. 75, Issue: 1-3, March, 1998, pp. 180-189.

[63] Yang, S.; Yuan, J. and Ni, J., “The improvement of thermal error modeling and

compensation on machine tools by CMAC neural network,” International Journal of

Machine Tools and Manufacture, Vol. 36, Issue: 4, April, 1996, pp. 527-537.

77
[64] Yang, S.; Yuan, J. and Ni, J., “Accuracy Enhancement of a Horizontal Machining

Center by Real-Time Error Compensation,” Journal of Manufacturing Systems, Vol.

15, Issue: 2, 1996, pp. 113-124.

[65] Yao, Y.; Li, J.; Lee, W.B.; Cheung, C.F. and Yuan, Z., “VMMC: a Test-Bed for

Machining,” Computers in Industry, Vol. 47, 2002, pp. 255-268.

[66] Zhang, H., Advanced Tolerancing Techniques, Chapter 20, John Wiley & Sons, 1997.

78

Вам также может понравиться