144
Contributions to Profile Monitoring and Multivariate Statistical Process Control James D. Williams Dissertation submitted to the faculty of the Virginia Polytechnic Institute & State University in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Statistics Jeffrey B. Birch, Co-Chairman William H. Woodall, Co-Chairman Christine M. Anderson-Cook Dan J. Spitzner G. Geoffrey Vining December 1, 2004 Blacksburg, Virginia KEYWORDS: Bioassay, False Alarm Rate, Functional Data, Heteroscedasticity, Hotelling’s T 2 Statistic, Lack-of-Fit, Minimum Volume Ellipsoid, Nonlinear Regres- sion, Sample Size, Successive Differences, Vertical Density Profile. c 2004 by James D. Williams ALL RIGHTS RESERVED

Contributions to Profile Monitoring and Multivariate ...€¦ · The content of this dissertation is divided into two main topics: 1) nonlinear profile monitoring and 2) an improved

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

  • Contributions to Profile Monitoring andMultivariate Statistical Process Control

    James D. Williams

    Dissertation submitted to the faculty of the

    Virginia Polytechnic Institute & State University

    in partial fulfillment of the requirements for the degree of

    Doctor of Philosophyin

    Statistics

    Jeffrey B. Birch, Co-Chairman

    William H. Woodall, Co-Chairman

    Christine M. Anderson-Cook

    Dan J. Spitzner

    G. Geoffrey Vining

    December 1, 2004

    Blacksburg, Virginia

    KEYWORDS: Bioassay, False Alarm Rate, Functional Data, Heteroscedasticity,

    Hotelling’s T 2 Statistic, Lack-of-Fit, Minimum Volume Ellipsoid, Nonlinear Regres-

    sion, Sample Size, Successive Differences, Vertical Density Profile.

    c© 2004 by James D. WilliamsALL RIGHTS RESERVED

  • Contributions to Profile Monitoring andMultivariate Statistical Process Control

    James D. Williams

    Abstract

    The content of this dissertation is divided into two main topics: 1) nonlinear profile

    monitoring and 2) an improved approximate distribution for the T 2 statistic based

    on the successive differences covariance matrix estimator.

    Nonlinear Profile Monitoring

    In an increasing number of cases the quality of a product or process cannot ad-

    equately be represented by the distribution of a univariate quality variable or the

    multivariate distribution of a vector of quality variables. Rather, a series of measure-

    ments are taken across some continuum, such as time or space, to create a profile. The

    profile determines the product quality at that sampling period. We propose Phase I

    methods to analyze profiles in a baseline dataset where the profiles can be modeled

    through either a parametric nonlinear regression function or a nonparametric regres-

    sion function. We illustrate our methods using data from Walker and Wright (2002)

    and from dose-response data from DuPont Crop Protection.

    Approximate Distribution of T 2

    Although the T 2 statistic based on the successive differences estimator has been

    shown to be effective in detecting a shift in the mean vector (Sullivan and Woodall

    (1996) and Vargas (2003)), the exact distribution of this statistic is unknown. An

    accurate upper control limit (UCL) for the T 2 chart based on this statistic depends on

    knowing its distribution. Two approximate distributions have been proposed in the

    literature. We demonstrate the inadequacy of these two approximations and derive

  • useful properties of this statistic. We give an improved approximate distribution and

    recommendations for its use.

    iii

  • Acknowledgments

    The first person to whom I owe an eternal debt of gratitude is my precious wife, Gina,

    who not only gave birth to two beautiful children since we moved to Blacksburg, but

    bore a disproportionate load of raising our three children while I worked towards

    finishing this degree. Since our marriage over five years ago, I have been a full-time

    graduate student. I am extremely thankful for how supportive she has been through

    these tough graduate school years.

    From the beginning I knew that Dr. Jeffrey B. Birch would not only be an

    inspirational teacher and mentor, but a good friend as well. After taking three classes

    from him, I was impressed with his ability to teach and inspire students to rise to

    their potential. I have tried to pattern my own teaching style according to his. Dr.

    Birch has been like a second father to me. He puts his own work on hold to hear my

    thoughts on a moments notice. I owe him a huge debt of gratitude for the countless

    selfless hours he spend teaching me, guiding me, counselling me, and simply listening

    to me. There are too many things to thank him for than can be adequately listed

    here.

    Dr. William H. Woodall has been an inspiration to me as well. At one point

    while deriving the results from Chapter 4, I asked him if he thought it would be

    alright if I used a simulation study to prove a theorem. His response was, “You

    could do that, but it would be better if you proved it analytically.” He left it at that,

    and I walked away scratching my head. It took me several months to figure it out,

    iv

  • but the analytical proof was finally completed. In addition to inspiring me to be a

    better researcher, Dr. Woodall helped me get through my final semester at Virginia

    Tech by selecting me to be supported in part by National Science Foundation Grant

    DMII-0354859.

    I also thank Dr. G. Geoffrey Vining for “that hallway conversation” during my

    first semester here at Virginia Tech, which lifted my sights and gave me a new vision

    of what I can become in the statistical profession. I also thank him for the many

    hours he spent preparing my dossier for the Virginia Tech College of Science Most

    Outstanding Graduate Student Award.

    During my tenure here at Virginia Tech, I counselled with many faculty and

    graduate students who greatly helped me. I thank Dr. George Terrell for always

    holding an “open door policy” and for giving many insightful hints that lead to

    big steps forward in completing my proofs. I thank Dr. J. P. Morgan for helping

    me to get started on my proofs. I also thank Mahmoud A. Mahmoud, Landon Sego,

    Willis Jensen, and Mike Joner for many insightful conversations in our quality control

    research group meetings.

    Most importantly, I lift a voice of gratitude to my Heavenly Father for hearing

    and answering my many sincere prayers for help in finishing this work. During the

    more difficult days I found myself on my knees multiple times in my graduate student

    carrell pleading for help. I acknowledge the hand of divinity in guiding my thoughts

    to find solutions when my mortal mind could not.

    — James D. Williams

    v

  • Contents

    List of Figures ix

    List of Tables xi

    Glossary of Acronyms xii

    Common Notation xiii

    1 Introduction 1

    1.1 Multivariate Statistical Process Control . . . . . . . . . . . . . . . . . 1

    1.1.1 Phase I Analysis . . . . . . . . . . . . . . . . . . . . . . . . . 1

    1.1.2 Phase II Analysis . . . . . . . . . . . . . . . . . . . . . . . . . 2

    1.1.3 Research Scope . . . . . . . . . . . . . . . . . . . . . . . . . . 3

    1.2 Nonlinear Profile Monitoring . . . . . . . . . . . . . . . . . . . . . . . 3

    1.3 Distribution of the T 2 Statistic . . . . . . . . . . . . . . . . . . . . . 4

    1.4 Example of Monitoring Dose-Response Profiles . . . . . . . . . . . . . 5

    1.5 Proposals for Future Work . . . . . . . . . . . . . . . . . . . . . . . . 5

    2 Literature Review 7

    2.1 Nonlinear Profile Monitoring . . . . . . . . . . . . . . . . . . . . . . . 7

    2.2 Distribution of the T 2 Statistic . . . . . . . . . . . . . . . . . . . . . 11

    vi

  • 3 Nonlinear Profile Monitoring 13

    3.1 Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

    3.1.1 Nonlinear Model Estimation . . . . . . . . . . . . . . . . . . . 14

    3.1.2 Multivariate T 2 Control Chart . . . . . . . . . . . . . . . . . . 16

    3.1.3 Control Limits . . . . . . . . . . . . . . . . . . . . . . . . . . 19

    3.1.4 Monitoring the Variance . . . . . . . . . . . . . . . . . . . . . 21

    3.1.5 Nonparametric Approach . . . . . . . . . . . . . . . . . . . . . 22

    3.2 Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

    3.3 Autocorrelation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

    3.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

    4 Distribution of the T 2 Statistic Based on the Successive Differences

    Covariance Matrix Estimator 37

    4.1 The T 2 Statistic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

    4.1.1 Asymptotic Marginal Distribution . . . . . . . . . . . . . . . . 39

    4.1.2 Approximate Marginal Distribution . . . . . . . . . . . . . . . 40

    4.2 Performance Comparison . . . . . . . . . . . . . . . . . . . . . . . . . 48

    4.2.1 Control Limits . . . . . . . . . . . . . . . . . . . . . . . . . . 48

    4.2.2 Simulation Study . . . . . . . . . . . . . . . . . . . . . . . . . 49

    4.3 Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

    4.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

    4.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

    5 Example of Monitoring Dose-Response Profiles from High Through-

    put Screening 56

    5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

    5.2 Bioassay Protocol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

    vii

  • 5.3 Methods: Homoscedastic Case . . . . . . . . . . . . . . . . . . . . . . 60

    5.3.1 Dose-Response Model . . . . . . . . . . . . . . . . . . . . . . . 60

    5.3.2 Phase I Profile Analysis . . . . . . . . . . . . . . . . . . . . . 61

    5.3.3 Phase II Profile Monitoring . . . . . . . . . . . . . . . . . . . 68

    5.4 Proposed Methods: Heteroscedastic Case . . . . . . . . . . . . . . . . 72

    5.4.1 Phase I Profile Analysis . . . . . . . . . . . . . . . . . . . . . 72

    5.4.2 Phase II Monitoring . . . . . . . . . . . . . . . . . . . . . . . 77

    5.5 Analysis of Dose-Response Profiles . . . . . . . . . . . . . . . . . . . 77

    5.5.1 Analysis Assuming Homoscedasticity . . . . . . . . . . . . . . 78

    5.5.2 Analysis Accounting for Heteroscedasticity . . . . . . . . . . . 85

    5.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

    6 Future Work and Conclusion 97

    6.1 Profile Monitoring . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

    6.2 Distribution of T 2D,i . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

    6.3 Combining Multivariate T 2 Control Charts . . . . . . . . . . . . . . . 100

    6.3.1 A Proposed Simulation Study . . . . . . . . . . . . . . . . . . 102

    6.3.2 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

    6.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

    A Appendix A: Result from Chapter 3 107

    B Appendix B: Results from Chapter 4 109

    B.1 Asymptotic Distribution of T 2D,i . . . . . . . . . . . . . . . . . . . . . 109

    B.2 Maximum Value of T 2D,i Statistics . . . . . . . . . . . . . . . . . . . . 110

    References 124

    Vita 130

    viii

  • List of Figures

    3.1 Vertical Density Profile (VDP) of 24 Particleboards . . . . . . . . . . 24

    3.2 “Bathtub” Function Fit to Board 1 . . . . . . . . . . . . . . . . . . . 26

    3.3 Nonlinear Regression Parameter Estimates a1, a2, b1, b2, c, and d by

    Board for the VDP Data . . . . . . . . . . . . . . . . . . . . . . . . . 28

    3.4 The T 2 Control Charts for the VDP Data . . . . . . . . . . . . . . . 29

    3.5 Spline Fit of Board 1 and Average Spline for the VDP Data . . . . . 31

    3.6 Control Charts on Metrics . . . . . . . . . . . . . . . . . . . . . . . . 32

    4.1 Q-Q plots of empirical quantiles of the T 2D,i statistic (i = 1 and 2)

    versus a χ2(p) distribution . . . . . . . . . . . . . . . . . . . . . . . . 40

    4.2 Boxplots of T 2D,i for p = 2 and m = 5. . . . . . . . . . . . . . . . . . . 42

    4.3 Q-Q plots of empirical quantiles of the scaled T 2D,i statistics for combi-

    nations of p = 4 and m = 30, 60 . . . . . . . . . . . . . . . . . . . . . 46

    4.4 Q-Q plots of empirical quantiles of the T 2D,i statistic for combinations

    of p = 8 and m = 30, 60 . . . . . . . . . . . . . . . . . . . . . . . . . 47

    4.5 Overall Probability of a False Alarm for p = 2, 3, 4, 5 . . . . . . . . . 50

    4.6 Overall Probability of a False Alarm for p = 6, 7, 8, 9 . . . . . . . . . 51

    4.7 T 2D,i statistics and UCL values for the Quesenberry (2001) data . . . . 53

    5.1 Estimated profiles for all 44 weeks, in a trellis plot. . . . . . . . . . . 80

    5.2 Estimated profiles for all 44 weeks, overlaid. . . . . . . . . . . . . . . 81

    ix

  • 5.3 Variance chart. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

    5.4 Lack-of-fit chart. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

    5.5 First T 2 chart for the mean profiles. . . . . . . . . . . . . . . . . . . . 83

    5.6 Second T 2 chart for the mean profiles. . . . . . . . . . . . . . . . . . 84

    5.7 Estimated in-control profiles. . . . . . . . . . . . . . . . . . . . . . . . 85

    5.8 Fitted variance profiles for all 44 weeks. . . . . . . . . . . . . . . . . . 87

    5.9 Fitted variance profiles for all 44 weeks, overlaid. . . . . . . . . . . . 88

    5.10 T 2 chart based on successive differences (a) and the MVE (b). . . . . 89

    5.11 Fitted mean profiles based on estimated weights for all 44 weeks. . . . 90

    5.12 Lack-of-fit chart based on the weighted sums of squares. . . . . . . . 91

    5.13 First T 2 chart for the mean profiles, heteroscedastic case . . . . . . . 91

    5.14 Second T 2 chart for the mean profiles, heteroscedastic case . . . . . . 92

    5.15 Estimated in-control mean profiles for the analysis assuming heteroscedas-

    ticity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

    x

  • List of Tables

    3.1 Estimated Parameter Values and T 2 Statistics for the VDP Data . . . 27

    4.1 The T 2D,i statistics scaled according to Sullivan and Woodall (1996),

    Mason and Young (2002), and Equation (4.6) for a data set. . . . . . 41

    4.2 The T 2D,i statistics and UCLvec values based on the Quesenberry (2001)

    data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

    5.1 Estimated parameter values, Wi, LOFi, and T2D,i statistics for the

    DuPont Crop Protection data . . . . . . . . . . . . . . . . . . . . . . 79

    5.2 S2ij values for every dose and week combination of the DuPont Crop

    Protection data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

    5.3 Estimated θ1,i (Slope) values, their standard errors, and 99.88% one-

    sided lower Wald confidence limit. . . . . . . . . . . . . . . . . . . . . 96

    xi

  • Glossary of Acronyms

    ANSS Average Number of Samples to Signal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

    ARL Average Run Length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

    ATS Average Time to Signal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

    GLIM Generalized Linear Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

    HDS Historical Dataset . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

    HTS High Throughput Screening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

    i.i.d. independent and identically distributed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

    IPP Inter-profile Pooling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

    LCL Lower Control Limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

    LOF Lack-of-fit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

    MCUSUM Multivariate Cumulative Sum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

    MEWMA Multivariate Exponentially Weighted Moving Average . . . . . . . . . . . . . . . 100

    MSE Mean Squared Error . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

    MVE Minimum Volume Ellipsoid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

    OD Optical Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

    PC Percent Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

    PE Pure Error . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

    POX Power of x . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

    SPC Statistical Process Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

    UCL Upper Control Limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

    VDP Vertical Density Profile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

    WNLS Weighted Nonlinear Least Squares . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

    xii

  • Common Notationyijk The response of replication k of the jth dose of profile i

    f(·) Nonlinear regression function for the meanX Historical dataset matrix

    βi Parameter vector for profile i

    Di Matrix of derivatives of f(·) with respect to βidi The number of doses in profile i (as in Chapter 5)

    rij The number of replications in dose j of profile i

    m The number of samples in the historical dataset

    p The number of parameters (dimension of βi)

    α The probability of a false alarm (for an individual observation)

    αoverall The probability that a control chart will have at least one false alarm

    θi Variance function parameter vector for profile i

    θ̂GLIM

    i Estimate of θi obtained by generalized linear models techniques

    S2ij Sample variance estimator for dose j in profile i

    SP Sample (pooled) variance-covariance matrix estimator

    SD Successive differences variance-covariance matrix estimator

    SIPP Intra-profile pooling variance-covariance matrix estimator

    SMV E Minimum volume ellipsoid variance-covariance matrix estimator

    T 2P,i T2i statistic based on SP

    T 2D,i T2i statistic based on SD

    T 2IPP,i T2i statistic based on SIPP

    T 2MV E,i T2i statistic based on SMV E

    T 2P T2 chart based on T 2P,i statistics

    T 2D T2 chart based on T 2D,i statistics

    T 2IPP T2 chart based on T 2IPP,i statistics

    T 2MV E T2 chart based on T 2MV E,i statistics

    MV (m, i) The maximum value of the T 2D,i statistic

    F(df1, df2) The F-distribution with df1 and df2 numerator and denominatordegrees of freedom, respectively

    F(q, df1, df2) The qth quantile of an F-distribution with df1 and df2 numeratorand denominator degrees of freedom, respectively

    BETA(p1, p2) The beta distribution with first and second shape parameters p1 andp2, respectively

    BETA(q, p1, p2) The qth quantile of a beta distribution with first and second shapeparameters p1 and p2, respectively

    χ2(p) The χ2 distribution with p degrees of freedom

    χ2(q, p) The qth quantile of a χ2 distribution with p degrees of freedom

    xiii

  • Chapter 1

    Introduction

    1.1 Multivariate Statistical Process Control

    Multivariate Statistical Process Control (SPC) is a broad field of research and ap-

    plications devoted to the improvement of products and processes. In monitoring

    the quality of a product or process, quite often more than one quality characteris-

    tic is measured on each manufactured item, thus producing a multivariate response.

    These quality measurements are usually correlated with each other. Multivariate

    SPC methods are designed to account for the correlations among the variables and

    to simultaneously monitor the variables through time.

    There are two phases of Multivariate SPC, Phase I and Phase II. A successful

    Phase II analysis depends on a successful Phase I analysis. Although the two phases

    are both dedicated to identifying out-of-control situations, each phase has a unique

    objective.

    1.1.1 Phase I Analysis

    The Phase I analysis begins with an analysis of a historical data set (HDS) or a set of

    baseline data consisting of multivariate observations taken on the product or process

    1

  • consecutively for a specified period of time. In a Phase I analysis one seeks to identify

    a stable subset of the HDS with which to estimate the in-control mean vector and the

    in-control variance-covariance matrix for use in a Phase II analysis. Some examples

    of unstable product or process conditions that one looks for are multivariate outliers,

    shifts in the mean vector, or shifts in the variance-covariance matrix. Once out-of-

    control data have been identified and eliminated, then a stable subset of the HDS is

    used to estimate in-control parameters.

    When evaluating the performance of competing control charts in a Phase I anal-

    ysis, one usually sets the false alarm rates of the competing charts to be the same

    by adjusting their respective control limits. Once the competing charts are set on

    “equal footing,” then one calculates the probability of signal for given shifts in the

    mean vector or covariance matrix, or some other out-of-control situation. The chart

    that has the highest probability of signal for a given shift is the most desirable chart.

    1.1.2 Phase II Analysis

    The Phase II analysis depends on the success of the Phase I analysis in estimating

    in-control mean, variance, and covariance parameters. The estimates of the in-control

    parameters become the target values in Phase II analysis control charts. The lower

    control limit (LCL) and upper control limit (UCL) also depend on the in-control

    parameter estimates. The purpose of a Phase II analysis is, thus, continuous process

    monitoring through time. Control limits are designed with the purpose of achieving

    an acceptable false alarm rate. If an out-of-control signal is found, then assignable

    causes of the signal are sought.

    When evaluating the performance of competing control charts in a Phase II analy-

    sis, one usually adjusts the control limits so that the in-control average time to signal

    (ATS) of the competing charts are equal. Other similar performance measures may

    2

  • also be used, such as the average run length (ARL) or the average number of samples

    to signal (ANSS). One then calculates the ATS (or other performance measure) value

    for a given shift from control for various out-of-control conditions. The chart that

    has the lowest ATS value is the desirable chart since the objective is quick detection

    of an out-of-control situation.

    1.1.3 Research Scope

    The scope of the research for this dissertation will be mostly restricted to Phase I

    analyzes. Correspondingly, the performance measure employed to evaluate competing

    control charts is the probability of signal given an out-of-control condition. In Chapter

    5 we give both Phase I and Phase II profile monitoring methods to monitor the dose-

    response quality profiles of herbicides in high throughput screening.

    This research is divided into two general areas: (1) profile monitoring and (2)

    properties of the multivariate T 2 control chart based on the successive differences

    variance-covariance matrix estimator. In Chapter 2 we review published literature on

    these two topics. Chapters 3 and 4 are dedicated to developing new statistical theory

    and methodology in these two areas. Chapter 5 contains an example of nonlinear

    profile monitoring applied to dose-response bioassay data. Chapter 6 contains a

    proposal for future work.

    1.2 Nonlinear Profile Monitoring

    In many quality control applications, use of a single (or several distinct) quality

    characteristic(s) is insufficient to characterize the quality of a produced item. In

    an increasing number of cases, a response curve (profile), is required. Such profiles

    can frequently be modeled using linear or nonlinear regression models. In recent

    research others have developed multivariate T 2 control charts and other methods for

    3

  • monitoring the coefficients in a simple linear regression model of a profile. However,

    little work has been done to address the monitoring of profiles that can be represented

    by a parametric nonlinear regression model.

    In Chapter 3 we extend the use of the T 2 control chart to monitor the coeffi-

    cients resulting from a nonlinear regression model fit to profile data. We give four

    general approaches to the formulation of the T 2 statistics and determination of the

    associated upper control limits for Phase I applications. We also consider the use of

    nonparametric regression methods and the use of metrics to measure deviations from

    a baseline profile. These approaches are illustrated using the vertical board density

    profile data presented in Walker and Wright (2002).

    1.3 Distribution of the T 2 Statistic

    In the historical or retrospective data analysis of Phase I, especially with individual

    observations, the choice of the estimator for the variance-covariance matrix is crucial

    to successfully detecting the presence of special causes of variation. The traditional

    estimator based on pooling all the historical observations is not useful in detecting a

    shift in the mean vector, because such a shift near the middle of the data actually

    reduces the probability that the corresponding T 2 chart will signal. The estimator

    based on successive differences is useful in detecting such shifts, but the exact dis-

    tribution for the corresponding T 2 chart statistic has not been determined. Three

    approximate marginal distributions have been proposed.

    In Chapter 4, several useful properties of the T 2 statistic based on the successive

    differences estimator are demonstrated and an improved distribution for calculating

    the UCL for individual observations in a Phase I analysis is given. This improved

    method for calculating the UCL is evaluated by comparing the estimated false alarm

    rate with the desired false alarm rate. It is shown that the proposed method achieves

    4

  • a false alarm rate which is closer to the desired false alarm rate than the competing

    methods when the sample size is small. For large sample sizes, the method based on

    the asymptotic distribution is shown to be superior. We give recommendations when

    to use each.

    1.4 Example of Monitoring Dose-Response Pro-

    files

    In pharmaceutical drug discovery and agricultural crop product development, in vitro

    bioassay experiments are used to identify promising compounds for future research.

    The reproducibility and accuracy of the bioassay is crucial to be able to correctly dis-

    tinguish between active and inactive compounds. In the case of agricultural product

    development, compound activity for a given test organism, such as weeds, insects, or

    fungi, is characterized by a dose-response curve measured from the bioassay. These

    curves determine the quality of the bioassay procedure. When undesirable conditions

    in the bioassay arise, such as equipment failure or hormetic effects, then a bioassay

    monitoring procedure is needed to quickly detect such causes. In Chapter 5 we illus-

    trate the proposed nonlinear profile monitoring methods to monitor the variability

    of assay, the adequacy of the dose-response model chosen, and the estimated dose-

    response curves for aberrant cases. We illustrate these methods with in vitro bioassay

    data from DuPont Crop Protection collected over one year.

    1.5 Proposals for Future Work

    The concept of combining two or more control charts into one overall monitoring

    procedure has been around for several years. The purpose of running two or more

    charts simultaneously is to leverage the strengths of each chart. For example, one

    5

  • chart may be sensitive to certain out-of-control situations and insensitive to others,

    whereas the second chart may have the opposite sensitivity properties. The idea

    is that the combination of the two charts will form an overall procedure which has

    sensitivity to both classes of out-of-control situations. It is conjectured that the chart

    combination’s loss in power to detect any specific out-of-control situation is small

    compared to the large gain in overall performance over a wider class of problems.

    As noted earlier, the purpose of a Phase I analysis is to identify a subset of stable

    data from the HDS with which to estimate the in-control mean vector and variance-

    covariance matrix. Two out-of-control situations that we seek to find are multivariate

    outliers and shifts in the mean vector. It has been shown by Vargas (2003) that the

    T 2 chart based on the minimum volume ellipsoid (MVE) mean and covariance matrix

    estimators is very effective in detecting multivariate outliers for a Phase I analysis.

    However, this chart is not very effective in detecting a shift in the mean vector.

    Sullivan and Woodall (1996) studied the performance of the T 2 chart based on the

    successive difference variance-covariance matrix estimator and found that this chart

    is very effective in detecting a shift in the mean vector. However, this chart is not

    very effective in detecting a multivariate outlier.

    In order to have a multivariate control chart that performs well in both detecting

    multivariate outliers and shifts in the mean vector, a new chart that combines the

    chart proposed by Vargas (2003) and Sullivan and Woodall (1996) could be explored.

    The details of how this chart is constructed are given in Chapter 6. The design for

    a simulation study is proposed to evaluate the power properties of this new chart

    compared to the Vargas (2003) chart alone, the Sullivan and Woodall (1996) chart

    alone, and the T 2 control chart based on the usual sample variance-covariance matrix

    estimator. The simulation consists of various out of control situations and the charts

    can be evaluated based on the probability of detecting the out-of-control observations.

    6

  • Chapter 2

    Literature Review

    2.1 Nonlinear Profile Monitoring

    In SPC applications, manufactured items are sampled over time and quality character-

    istics are measured. Often a product’s quality can be determined through measuring

    several characteristics at each sampling interval. Multivariate T 2 control charts and

    other methods have been developed for this scenario. See, for example, Fuchs and

    Kenett (1998) and Mason and Young (2002). Increasingly, however, a sequence of

    measurements of one or more quality characteristics are taken across some continuum

    producing a curve or surface that represents the quality of the item. This curve or

    surface is referred to as a profile. Very little work has been done in developing sta-

    tistical process control methodology for monitoring profile data. For an overview of

    profile monitoring techniques see Woodall, Spitzner, Montgomery, and Gupta (2004).

    Profile data consist of a set of measurements with a response variable y and one

    or more explanatory variables xj, j = 1, . . . , k, which are used to assess the quality of

    a manufactured item. For example, the density profile of a particleboard is measured

    on a vertical cross-section, which reveals patterns in board density across the depth of

    the board. Another example is the estimated dose-response curve of a manufactured

    7

  • drug. Once a batch of the drug is produced, several different doses of the drug are

    administered to subjects and the responses measured. The resultant dose-response

    curve summarizes the quality of the particular batch of the drug, indicating the maxi-

    mal effective response, minimal effective response, and the rate in which the response

    changes between the two. In these examples, a single measurement is insufficient to

    adequately assess quality. Instead, a relationship between two variables, referred to

    as the profile, should be monitored over time. Profile data is multivariate, but it is

    not appropriate to apply standard multivariate control chart methods since this leads

    to overparameterization. It is more efficient to model the structure of the data.

    Profiles can take on several different functional forms, depending on the specific

    application. For many calibration problems, the profile can be represented by a

    simple linear regression model (see, e.g., Mahmoud and Woodall (2004)). Kang and

    Albin (2000) proposed two methods, including a multivariate T 2 control chart, to

    monitor such profiles. Specifically, we let the subscript i index each individual profile

    (i = 1, . . . , m) in the historical Phase I data. In the simple linear regression case, the

    ith profile is modeled as

    yij = βi0 + βi1xij + ²ij, (2.1)

    where yij is the jth measurement (j = 1, . . . , n), ²ij is the j

    th random error, and xij is

    the jth value of the explanatory variable corresponding to the ith profile. It is assumed

    that the values of xij are the same for all i. This assumption is often reasonable since

    in many engineering applications product or process profiles are measured at fixed

    values of the explanatory variable at each sampling stage. Kang and Albin’s (2000)

    multivariate T 2 chart is used to monitor simultaneously the β0, the y-intercept, and

    β1, the slope. Kim, Mahmoud, and Woodall (2003) proposed an alternative approach

    with better statistical properties such that individual control charts can be used for

    the y-intercept and slope independently.

    8

  • In general we refer to any profile that can be modeled by the linear regression

    function

    yij = βi0 + βi1xij1 + βi2xij2 + · · ·+ βikxijk + ²ij (2.2)

    as a linear profile, where xijl, l = 1, . . . , k, are k predictor variables. The predictor

    variables can be the original variables themselves, any function of the variables, or

    any combination of both. In matrix notation, we let yi = [yi1, yi2, . . . , yin]′ be the

    vector of responses for profile i, βi = [βi0, βi1, . . . , βik]′ be the vector of parameters to

    be monitored, x′ij = [1, xij1, xij2, . . . , xijk] be the vector of explanatory variables for

    item i, and ²i = [²i1, ²i2, . . . , ²in]′ be the corresponding vector of random errors. After

    collecting the x′ij vectors into an n× p matrix, where p = k + 1, as

    Xi =

    x′i1x′i2...

    x′in

    ,

    then model (2.2) can be written in matrix form as

    yi = Xiβi + ²i, i = 1, . . . ,m.

    We assume that Xi is the same for each profile and that the vectors ²i are independent

    and identically distributed (i.i.d.) multivariate normal random vectors with mean

    vector zero and covariance matrix σ2I.

    Jensen, Hui, and Ghare (1984) proposed a control chart based on the F -distribution

    to monitor the k +1 parameters (coefficients) from a multiple linear regression model

    for Phase II applications. Given the parameter vector estimator for item i, β̂i, and

    the target parameter vector β0, one plots on their control chart the well-known F

    statistic

    Fi = (β̂i − β0)′X′iXi(β̂i − β0)/(k + 1)s2i

    against i, where s2i =∑n

    i=1(yi − ŷ)2/(n − p). A Phase I procedure for this generallinear case has yet to be developed.

    9

  • In many cases, however, profiles cannot be well-modeled by a linear regression

    function. Walker and Wright (2002) proposed a nonparametric approach for com-

    paring profiles using additive models. Such models do not have a specific functional

    form and have no model parameters to estimate, but rather one employs smooth-

    ing techniques such as local polynomial regression or spline smoothing to model a

    profile. Nonparametric regression techniques provide great flexibility in modeling the

    response. One disadvantage of nonparametric smoothing methods is that the subject-

    specific interpretation of the estimated nonparametric curve may be more difficult,

    and may not lead the user to discover as easily assignable causes that lead to an

    out-of-control signal.

    Often, however, scientific theory or subject-matter knowledge leads to a natural

    nonlinear function that well-describes the profiles. Hence, an alternative method is

    to model each profile by a nonlinear regression function. A nonlinear profile of an

    item can be modeled by the nonlinear regression model given generally by

    yij = f(xij,βi) + ²ij, (2.3)

    where xij is a k× 1 vector of regressors for the jth observation of the ith profile, ²ij isthe random error, βi is a p×1 vector of parameters for profile i, and f(·) is a functionwhich is nonlinear in the parameter vector βi. The random errors ²ij are assumed

    to be i.i.d. normal random variables with mean zero and variance σ2. In many

    applications, there is only one regressor (k = 1), but there are multiple parameters

    to monitor (p > 1). An example of this form of the model is the 4-parameter logistic

    model, often used to model dose-response profiles of a drug, given by

    yij = Ai +Di − Ai

    1 +

    (xijCi

    )Bi + ²ij, (2.4)

    where yij is the measured response of the subject exposed to dose xij for batch i,

    i = 1, . . . , m, j = 1, . . . , d, where d is the number of doses. In equation (2.4), we

    10

  • have k = 1 and p = 4, giving four parameters to monitor, each parameter having a

    specific interpretation. For example, Ai is the upper asymptote parameter, Di is the

    lower asymptote parameter, Ci is the ED50 parameter (the dose required to elicit a

    50% response), and Bi is the rate parameter for the ith batch. Another example is

    the “bathtub” function described in Section 3.2 where the density of particleboard is

    measured across the vertical profile. Note that for any given application, the specific

    form of the nonlinear function, f , in equation (2.3) must be specified by the user.

    In Phase I analysis, we are concerned with distinguishing between in-control con-

    ditions and the presence of assignable causes so that in-control parameters may be

    estimated for further product or process monitoring in Phase II analysis. In Chapter

    3 we discuss some procedures for Phase I analysis for monitoring items or processes

    whose quality is reflected by a nonlinear profile.

    2.2 Distribution of the T 2 Statistic

    Multivariate SPC is prevalent in many aspects of industry and manufacturing wher-

    ever several different measures of a product or process are taken at each sampling

    stage to assess quality. A common statistical method used to simultaneously monitor

    the multiple quality characteristics is use of the Hotelling T 2 statistic.

    For a retrospective Phase I analysis of an HDS the objective is twofold: (1) to

    identify and eliminate multivariate outliers, and (2) to identify shifts in the mean

    vector which might distort the estimation of the in-control mean vector and variance-

    covariance matrix. The T 2 control chart is a tool to detect multivariate outliers and

    mean shifts. The T 2 statistic one plots in the chart can be based on the usual sample

    variance-covariance matrix estimator or some other alternative. One alternative is

    the estimator based on the successive differences of observation vectors. As shown in

    both Sullivan and Woodall (1996) and Vargas (2003) the T 2 statistic based on the

    11

  • successive differences variance-covariance matrix estimator is effective in detecting

    sustained step and ramp shifts in the mean vector. Sullivan and Woodall (1996)

    found that the T 2 statistic based on the usual sample variance-covariance matrix

    estimator is not only less effective in detecting a shift in the mean vector, but as the

    magnitude of the shift increased, the power to detect the shift decreased. They found

    that the sample variance-covariance matrix has the effect of “pooling” the data all

    together such that a large step shift “inflates” the variance, thus making detection of

    the shift more difficult.

    Imperative to constructing any multivariate control chart is knowing the marginal

    distribution of the test statistic. The UCL of the control chart is calculated from a

    specified quantile of this marginal distribution. If the marginal distribution of the

    test statistic is unknown or untractable, then the UCL is calculated from either an

    approximate marginal distribution, where available, or from a Monte Carlo simula-

    tion. If an approximate marginal distribution is used, the researcher should be aware

    of the cases under which the approximation performs well and when it does not.

    Unfortunately, the exact small-sample marginal distribution of the T 2 statistic

    based on the successive differences variance-covariance matrix estimator is unknown.

    Two approximate marginal distributions have been proposed, one by Sullivan and

    Woodall (1996) and the other by Mason and Young (2002). Another possible approx-

    imate marginal distribution is the asymptotic distribution. In Chapter 4 we give the

    asymptotic marginal distribution and give recommendations for its use. We propose

    an improved small-sample marginal distribution and demonstrate that the proposed

    approximate distribution gives rise to UCLs that perform much better than the other

    approaches for small sample sizes. We will discuss some useful properties of the dis-

    tribution of the T 2 statistic based on the successive differences variance-covariance

    matrix estimator and compare the performance of the two approximate distributions.

    12

  • Chapter 3

    Nonlinear Profile Monitoring

    In Section 3.1 we give a brief review of nonlinear regression. We introduce the multi-

    variate T 2 statistic in the context of monitoring nonlinear profiles. We then introduce

    four formulations of the T 2 statistic and discuss the determination of the UCL for the

    corresponding charts. In addition, a control chart to monitor the variance σ2 in the

    context of monitoring profile data is proposed. Finally, we discuss a nonparametric

    regression approach to monitoring the profiles. In Section 3.2 we illustrate the T 2

    control charts and the nonparametric approaches using the vertical density profile

    data of Walker and Wright (2002). In Section 3.3 we discuss the effects that auto-

    correlation in the error terms may have on the analysis. Finally, in Section 3.4 we

    discuss potential alternative methods and give directions for future research topics in

    nonlinear profile monitoring.

    3.1 Methodology

    We begin a Phase I analysis with a baseline dataset consisting of m items sampled

    over time. For each item i we observe a response yij and a set of predictor vari-

    ables xij, i = 1, . . . ,m, j = 1, . . . , n, resulting in the quality profile for item i, i.e.,

    (yi1,xi1), (yi2,xi2), . . . , (yin,xin). In this section we develop the methodology to ana-

    lyze the profiles to gain understanding of the product or process in a Phase I setting.

    13

  • 3.1.1 Nonlinear Model Estimation

    For simplicity of notation, we write the scalar model given in equation (2.3) in matrix

    form by stacking the n observations within each profile as yi = (yi1, yi2, . . . , yin)′,

    f(Xi, βi) = (f(xi1,βi), f(xi2,βi), . . . , f(xin,βi))′, and ²i = (²i1,²i2, . . . ,²in)′. The vec-

    tor form is then given by

    yi = f(Xi,βi) + ²i, i = 1, . . . , m. (3.1)

    For the nonlinear regression model given in (3.1), estimates of βi for each sample

    must be obtained. This is usually accomplished by employing the Gauss-Newton pro-

    cedure and iterating until convergence to obtain the maximum likelihood estimates.

    Specifically, for each sampling stage we define the m × p matrix of derivatives off(Xi,βi) with respect to βi as

    Di =∂f(Xi, βi)

    ∂βi=

    ∂f(xi1,βi)∂βi1

    ∂f(xi1,βi)∂βi2

    . . . ∂f(xi1,βi)∂βip

    ∂f(xi2,βi)∂βi1

    ∂f(xi2,βi)∂βi2

    . . . ∂f(xi2,βi)∂βip

    ......

    . . ....

    ∂f(xin,βi)∂βi1

    ∂f(xin,βi)∂βi2

    . . . ∂f(xin,βi)∂βip

    . (3.2)

    We let f(Xi, β̂

    (h)

    i

    )=

    (f(xi1, β̂

    (h)

    i ), f(xi2, β̂(h)

    i ), . . . , f(xin, β̂(h)

    i ))′

    , where β̂(h)

    i is the

    estimate of β at iteration h and we let D̂(h)i be the matrix of derivatives of f given

    in equation (3.2) evaluated at β̂(h)

    i . Then an iterative solution for β̂i is given by

    β̂(h+1)

    i = β̂(h)

    i +(D̂i

    ′(h)D̂

    (h)i

    )−1D̂′(h)i

    (yi − f(Xi, β̂(h)i )

    ).

    Upon convergence of the algorithm, the estimated covariance matrix of β̂i is the

    estimated Fisher information matrix and is given as

    ˆV ar(β̂i

    )= σ̂2i (D̂

    ′iD̂i)

    −1,

    where σ̂2i =∑m

    j=1(yij−f(xij, β̂i))2/(n−p) and D̂i is the derivative matrix in equation(3.2) evaluated at the converged parameter vector estimate β̂i. See Myers (1990, chap.

    14

  • 9) or Schabenberger and Pierce (2002, chap. 5) for a concise discussion of nonlinear

    regression model estimation. A more detailed treatment can be found in Gallant

    (1987) and Seber and Wild (1989).

    Unlike linear regression, the small-sample distribution of parameter estimators in

    nonlinear regression is unobtainable, even if the errors ²ij are assumed to be i.i.d.

    normal random variables. Instead, asymptotic results must be applied. Seber and

    Wild (1989, chap. 12) give the asymptotic distribution of β̂i and the necessary

    assumptions and regularity conditions for the asymptotic distribution to be obtained.

    Given their assumptions and assuming that n−1D̂′iD̂i converges to some nonsingular

    matrix Ωi as n →∞, the asymptotic distribution of β̂i is given by

    √n(β̂i − βi) D−→ Np(0, σ2Ω−1i ) (3.3)

    where βi is the asymptotic expected value of β̂i and Np indicates a p-dimensional mul-

    tivariate normal distribution. For practical purposes, the distribution given by (3.3)

    is incalculable since the matrix Ωi is unknown. Instead the approximate asymptotic

    distribution of β̂i is commonly used, given by

    β̂i·∼ Np(βi, σ2i (D′iDi)−1). (3.4)

    The standard estimator of V ar(β̂i) is σ̂2i D̂

    ′iD̂i. For the most traditional “in-control”

    case, we have βi = β for all profiles i = 1, . . . , m, where β is the in-control parameter

    vector. Consequently, the Ωi and Di matrices are the same across all m items since

    all items have the same underlying model, f , the same x-values are observed, and the

    same values of βi. However, the D̂i matrices are not equal since the β̂i values vary

    from profile to profile.

    15

  • 3.1.2 Multivariate T 2 Control Chart

    In order to develop the methodology to monitor nonlinear profiles, we first consider the

    general framework of the multivariate T 2 statistic. Given a sample of m independent

    observation vectors to be monitored, wi (i = 1, . . . ,m), each of dimension p, the

    general form of the T 2 statistic in Phase I for observation i is

    T 2i = (wi − w̄)′ S−1 (wi − w̄) , (3.5)

    where w̄ = 1m

    ∑mj=1 wi and S is some estimator of the variance-covariance matrix of

    wi (Mason and Young 2002, chap. 2). We then plot the T2i statistics, i = 1, . . . ,m,

    against i in a T 2 control chart, and out-of-control signals will be given for any T 2i value

    exceeding the UCL. For determining the statistical properties of the T 2 chart it is

    usually assumed that each of the wi vectors follows a multivariate normal distribution

    with common mean vector µ and covariance matrix Σ. This assumption is critical

    to finding the marginal distribution of T 2i , as discussed in Section 3.1.3.

    In the nonlinear regression model given in equation (2.3), βi is a p × 1 vectorof parameters that determines the curve f(Xi,βi). We employ the multivariate T

    2

    statistic to assess stability of the the p parameters simultaneously, i.e., to evaluate

    the assumption βi = β, i = 1, . . . , m. We do not employ individual control charts for

    each of the p nonlinear regression parameters since this may give misleading results

    due to the built-in correlation structure of the parameter estimators in nonlinear

    regression.

    Once β̂i is obtained from each sample in the baseline dataset, we calculate the

    average vector¯̂β and some corresponding estimate of the covariance matrix, replace

    wi with β̂i and w̄ with¯̂β in equation (3.5) to obtain

    T 2i =(β̂i − ¯̂β

    )′S−1

    (β̂i − ¯̂β

    ). (3.6)

    A large value of T 2i indicates an unusual β̂i, suggesting that the profile for item i

    16

  • is out-of-control. In contrast to the traditional use of the T 2 statistic to monitor a

    multivariate quality characteristic vector, we employ the T 2 statistic to monitor the

    coefficient vectors of the nonlinear regression fit to each individual profile.

    There are several choices for the estimator S. Here we discuss the effects of four

    choices and later discuss under what conditions, if any, each should be used.

    The first choice we consider for S is the sample covariance matrix, given by

    SP =1

    m− 1m∑

    i=1

    (β̂i − ¯̂β

    )(β̂i − ¯̂β

    )′. (3.7)

    Consequently, the T 2i statistics take on the form

    T 2P,i =(β̂i − ¯̂β

    )′S−1P

    (β̂i − ¯̂β

    ).

    We use the subscript “P ” to emphasize the fact that SP is the sample variance-

    covariance matrix based on pooling all the data in the HDS. We refer to the T 2 chart

    based on values of T 2P,i as the T2P chart. Use of the T

    2P,i values was mentioned by

    Brill (2001) in the context of monitoring nonlinear profiles of a chemical product.

    The advantage of this statistic is that it is very well understood and widely used.

    However, as was shown by Sullivan and Woodall (1996) and Vargas (2003), a T 2

    statistic based on SP is ineffective in detecting sustained shifts in the mean vector

    during the Phase I period. In fact, it was shown that as the step shift size increased,

    the power to detect the shift actually decreased.

    An alternative choice of S is one based on successive differences, proposed origi-

    nally by Hawkins and Merriam (1974) and later by Holmes and Mergen (1993). To

    obtain the estimator, we define vi = β̂i+1 − β̂i for i = 1, . . . , m − 1 and stack thetranspose of these m− 1 difference vectors into the matrix V as

    V =

    v′1v′2...

    v′m−1

    .

    17

  • The estimator of the variance-covariance matrix is

    SD =V′V

    2(m− 1) . (3.8)

    Sullivan and Woodall (1996) showed that SD is an unbiased estimator of the true

    covariance matrix if the process is stable in Phase I. The resulting T 2i statistics are

    given by

    T 2D,i =(β̂i − ¯̂β

    )′S−1D

    (β̂i − ¯̂β

    ). (3.9)

    We refer to the T 2 chart based on values of T 2D,i as the T2D chart. Sullivan and Woodall

    (1996) showed that the T 2D chart was effective in detecting both a step and ramp shift

    in the mean vector during Phase I. They also showed that the T 2D,i values are invariant

    to a full-rank linear transformation on the observations.

    Neither SP or SD, however, directly incorporate the information on the variation

    in the regression parameter estimators from the nonlinear regression estimation of

    β̂i. Consequently, a third choice for S is to pool the m covariance matrices resulting

    from the estimation of each β̂i vector. We refer to this method of calculating S as

    intra-profile pooling (IPP). If we let D̂i be the derivative matrix for item i evaluated

    at converged parameter vector estimate β̂i, then a third estimator of the variance-

    covariance matrix is

    SIPP =1

    m

    m∑i=1

    ˆV ar(β̂i

    )=

    1

    m

    m∑i=1

    σ̂2i (D̂′iD̂i)

    −1. (3.10)

    Consequently, the T 2 statistics are given as

    T 2IPP,i =(β̂i − ¯̂β

    )′S−1IPP

    (β̂i − ¯̂β

    ), i = 1, . . . , m. (3.11)

    We refer to the T 2 chart based on values of T 2IPP,i as the T2IPP chart.

    Our fourth choice for S is a robust estimator of the variance-covariance ma-

    trix known as the minimum volume ellipsoid (MVE) estimator, first proposed by

    18

  • Rousseeuw (1984). In our application of the MVE method, we find outlier-robust es-

    timates for both the in-control parameter vector and the variance-covariance matrix

    based on finding the ellipsoid with the smallest volume that contains at least half

    of the β̂i vectors, i = 1, . . . , m. The MVE estimator of β is the mean vector of the

    smallest ellipsoid, and the estimator of the variance-covariance matrix is the sample

    covariance matrix of the observations within the smallest ellipsoid multiplied by a

    constant to make the estimator unbiased for multivariate normal data. In a simu-

    lation study, Vargas (2003) studied the power properties of several different choices

    of S in the context of the T 2 statistic given in (3.5) and found that the T 2 statistic

    based on the MVE estimators of β and the variance-covariance matrix is very pow-

    erful in detecting multivariate outliers. We denote the MVE estimators of β and the

    covariance matrix by β̂MV E and SMV E, respectively. Hence, the fourth choice of T2

    is

    T 2MV E,i =(β̂i − β̂MV E

    )′S−1MV E

    (β̂i − β̂MV E

    ), i = 1, . . . , m. (3.12)

    We refer to the T 2 chart based on values of T 2MV E,i as the T2MV E chart.

    3.1.3 Control Limits

    The distribution of the T 2i statistics for monitoring nonlinear profiles is more complex

    than in the linear profile case. Recall that the distribution of the parameter estimators

    in nonlinear regression is difficult to obtain for small sample sizes. Instead we employ

    the asymptotic distribution (as n → ∞) of β̂i, i = 1, . . . , m. Hence, in order todetermine the marginal distribution of T 2i in this case, we assume that the sample size,

    n, from each item in the baseline data set is of sufficient size such that the distributions

    of β̂i, i = 1, . . . , m are approximately multivariate normal. The subsequent UCLs for

    the multivariate T 2 control charts are determined based on this normality assumption.

    In order to control the overall probability of a false alarm, based on some ap-

    19

  • propriate UCL, the joint distribution of the T 2i values is required. However, these

    values are correlated since¯̂β and S are used in all T 2i statistics (i = 1, . . . , m), thus

    making the joint distribution of the T 2i values difficult to obtain. As an alternative,

    Mahmoud and Woodall (2004) suggested using an approximate joint distribution as-

    suming that the T 2i statistics are independent. We let α be the probability of a false

    alarm for any individual T 2i statistic. Then the approximate overall probability of a

    false alarm for a sample of m items is given by αoverall = 1 − (1 − α)m. Thus, fora given overall probability of a false alarm, we use α = 1 − (1 − αoverall)1/m in thecalculation of UCLs. In their simulation study, Mahmoud and Woodall (2004) found

    that this approximation used to determine the UCLs performed well.

    As noted in Tracy, Young, and Mason (1992), Gnanadesikan and Kettenring

    (1972) suggested that for a stable process the marginal distribution of the T 2P,i statistic

    is proportional to a beta distribution, i.e.,

    T 2P,im

    (m− 1)2 ∼ BETA(

    p

    2,m− p− 1

    2

    ).

    A formal proof can be found in Chou, Mason, and Young (1999). Note that it

    is assumed that the distribution of β̂i is approximately normal, as given in equa-

    tion (3.4). Therefore, an approximate UCL is (m−1)2

    mBETA(1 − α, p

    2, m−p−1

    2), where

    BETA(1 − α, p2, m−p−1

    2) is the 1 − α quantile of a beta distribution with first and

    second shape parameters p/2 and (m− p− 1)/2, respectively.Sullivan and Woodall (1996) proposed an approximate marginal distribution for

    the T 2D,i statistic as

    T 2D,im

    (m− 1)2 ∼ BETA(

    p

    2,f − p− 1

    2

    ), (3.13)

    where f = 2(m−1)2

    3m−4 . Mason and Young (2002, pp. 26-27) gave a correction to the

    distribution proposed by Sullivan and Woodall (1996), by replacing each m in equa-

    tion (4.4) with f . Hence, an approximate UCL is (f−1)2

    fBETA(1− α, p

    2, f−p−1

    2). Our

    20

  • simulation results, not shown here, showed that neither one of these approximations

    worked well in obtaining the appropriate UCL, so in our example in Section 3.2 we

    used simulation to find the appropriate UCL. In Chapter 4 we give an approximate

    distribution that yields improved UCL values.

    Since the small-sample exact distribution of the T 2IPP,i statistic is difficult to obtain

    we use instead the asymptotic distribution. As proven in Appendix A, the asymptotic

    distribution is χ2(p). The asymptotic distribution of T 2IPP,i must be used in practice

    since the small-sample distribution is unknown. Hence, an approximate UCL is χ21−α,p.

    The exact marginal distribution of the T 2MV E,i statistic is unknown and intractable.

    Hence, in order to find the UCL for T 2MV E,i we used simulation.

    3.1.4 Monitoring the Variance

    In addition to checking the stability of each profile in the baseline dataset, it is

    important to check the stability of the variability about each profile. This is anal-

    ogous to monitoring the process variance in the standard univariate case. In the

    case of monitoring profiles, we seek to monitor the variability about each profile, or

    the within-profile variability. Our measure of within-profile variability is the mean

    squared error (MSE) defined as MSEi =∑n

    j=1(yij − ŷij)2/(n − p), where ŷij is thepredicted value of yij based on the nonlinear regression model in equation (2.3).

    Wludyka and Nelson (1997) recommended a method to monitor variances based on

    an analysis-of-means-type test utilizing S2i = MSEi. In their paper, S2i is plotted

    against i with associated lower- and upper-control limits equal to (Lα,m,n−p)mS2 and

    (Uα,m,n−p)mS2, respectively, where L and U are critical values given in their paper

    and S2 is the average of the S2i values, i = 1, . . . , m. For large n, their approximate

    upper and lower control limits are S2 ± hα,m,∞σ̂ where h is a critical value given inNelson (1983) and σ̂ = S2

    √2(m− 1)/m(n− p). The S2i statistics are plotted on

    21

  • a separate control chart to monitor the variance of the error terms and lack of fit

    simultaneously with a T 2 control chart for the nonlinear regression parameters. We

    recommend use of this method when within-profile error terms are independent. An

    approach to monitor the variance when there is replication is given in Chapter 5.

    3.1.5 Nonparametric Approach

    When a parametric form of a profile would be overly complex, nonparametric proce-

    dures may be more appropriate. These include fitting each profile via some smoothing

    method, such as local polynomial regression, spline smoothing, or wavelets. Walker

    and Wright (2002) give a spline-fitting approach to the vertical density profile (VDP)

    of particleboard, which we use as an illustration in Section 3.2. However, these au-

    thors discussed using splines to assess variation, not to monitor profiles in a Phase I

    analysis to check for process stability. Winistorfer, Young, and Walker (1996) illus-

    trated the use of splines to model the VDP of oriented strandboard generated from

    a 32 factorial design with 3 replicates. However, their spline-fitting method is used

    in the context of comparing profiles among differing experimental conditions, not

    monitoring profiles in a Phase I or Phase II analysis.

    For the case of a single explanatory variable, we denote the nonparametric fit of

    profile i by ẏij, for the corresponding value of the explanatory variable equal to xj,

    j = 1, . . . , n. The general nonparametric approach to monitoring profiles in Phase I

    analysis is to establish a “baseline” curve with which to compare all other curves. A

    natural choice of baseline profile is the average estimated profile across all m profiles,

    denoted by ỹj =∑m

    i=1 ẏij/m, j = 1, . . . , n. Once a baseline curve is found, some

    appropriate distance metric can be used to measure how “different” each individual

    curve is from the baseline. Researchers at Boeing (1998, pp. 140-144) proposed the

    following three metrics:

    22

  • 1. Mi1 = sign(ẏij∗ − ỹij∗)×maxj |ẏij − ỹij|2. Mi2 =

    ∑nj=1 |ẏij − ỹij|

    3. Mi3 =∑n

    j=1 |ẏij − ỹij|/m,

    where j∗ is the value of j that produces the maximum absolute deviation. The three

    metrics, Mi1, Mi2, and Mi3, are referred to as the maximum deviation, sum of absolute

    deviations, and the mean absolute deviation, respectively. Further, it may be of

    interest to compute the absolute value of Mi1, which obviously reflects the magnitude

    of the dissimilarity between ẏij and ỹij disregarding the direction of dissimilarity. We

    denote this metric by Mi4. Other metrics are proposed in Gardner, et. al. (1997), who

    note that metrics can be defined to detect changes in profiles resulting from particular

    known process faults. One of these metrics is the sum of squared differences between

    each estimated profile and the average profile, denoted Mi5 =∑n

    j=1(ẏij − ỹij)2. Fora given metric, one plots the metric value for profile i against i (i = 1, . . . , m) and

    checks for unusual observations. Researchers at Boeing (1998) suggested using a

    standard univariate I-chart on the metrics to establish control limits. The method of

    smoothing splines with several dissimilarity metrics is illustrated in Section 3.2.

    3.2 Example

    To illustrate the application of the various approaches we use the vertical density pro-

    file data from Walker and Wright (2002), available at the website http://bus.utk.edu/

    stat/walker/VDP.Allstack.TXT. In the manufacture of particleboard, the density

    properties of the finished boards are quality characteristics that are monitored through

    time. It is well known that the density (in pounds per ft3) near the core of a parti-

    cleboard is much less than the density at the top and bottom faces of a board (see

    Young, Winistorfer, and Wang (1999)). The standard sampling procedure calls for

    a laser-aided density measuring device to scan fixed vertical depths of a board and

    23

  • record the density at each depth. Since the depths are fixed for each board, we de-

    note the depth xij by simply xj. Density measurements for this dataset were taken

    at depths of xj = (0.002)j inches, j = 0, 1, 2, . . . , 313. Correspondingly, a sequence

    of ordered pairs, (xi, yij), j = 1, . . . , n, results for board i and forms a vertical density

    profile (VDP) of the board. A baseline sample of twenty-four particleboards was

    measured in this way, and the twenty-four profiles are illustrated in Figure 3.1.

    Figure 3.1: Vertical Density Profile (VDP) of 24 Particleboards

    0.0 0.1 0.2 0.3 0.4 0.5 0.6

    3540

    4550

    5560

    65

    Depth (inches)

    Den

    sity

    (lbs

    ft3)

    An approach to understand the profile variation in Phase I is a method proposed

    by Jones and Rice (1992) and discussed in Woodall, et. al. (2004). This method is

    based on principal components where each profile is represented as an n × 1 vectorof responses and mutually orthogonal linear combinations of the responses are found

    which explain as much variation as possible. Jones and Rice (1992) then recommended

    24

  • plotting the profiles with the largest and smallest principal component scores to give

    a visual interpretation of what each principal component represents. An example of

    this method using the VDP data can be found in Woodall, et. al. (2004). They show

    that the first two principal components correspond to the level and the flatness of

    the profiles and explain 84% and 10.77% of the variation, respectively. We strongly

    recommend the use of such plots.

    Young, Winistorfer, and Wang (1999) introduced a statistical method to monitor

    VDP data. With their method, one summarizes the density measurements into three

    average density measurements: one near the core and one near each face. The three

    averages are the quality characteristics that are subsequently monitored using a stan-

    dard multivariate T 2 control chart. With this method one basically summarizes each

    nonlinear profile into only three numbers with a corresponding loss of information.

    An alternative approach without such a considerable loss of information is to

    model the profiles themselves parametrically. The nonlinear function we use to model

    profile i is a “bathtub” function given by

    f(xij,β) =

    a1(xij − c)b1 + d xj > ca2(−xij + c)b2 + d xj ≤ c

    i = 1, . . . ,m; j = 1, . . . , n, (3.14)

    where β = (a1, a2, b1, b2, c, d)′. One advantage of this nonlinear model is the inter-

    pretability of the model parameters. For example, a1, a2, b1, and b2 determine the

    “flatness”, c is the center, and d is the bottom, or the “level” of the curve. Differing

    values of a1 and a2 or different values of b1 and b2 allow for an asymmetric curve about

    the center c. Other parameterizations of this model are possible. However, reparam-

    eterizing can effect the outcome of the Phase I analysis since different parameters

    can produce different T 2 statistics. A strategy one should take is to parameterize

    the model so that parameters of interest can be estimated and monitored. For this

    example, the proposed parameterization addresses the shape, center, and symmetry

    25

  • of the profiles. Figure 3.2 contains the “bathtub” function fit to board 1 from the

    VDP data.

    Figure 3.2: “Bathtub” Function Fit to Board 1

    0.0 0.1 0.2 0.3 0.4 0.5 0.6

    4045

    5055

    6065

    Depth (inches)

    Den

    sity

    (lbs

    ft3)

    The profile of board 1 is well-modeled by this parametric fit (R2 > 0.9999). For

    each of the twenty-four boards in the baseline sample we fit the nonlinear model

    in equation (3.14), and calculated the T 2P,i, T2D,i, T

    2IPP,i, and T

    2MV E,i statistics based

    on the β̂i values. Parameter estimates for each of the twenty-four boards and the

    corresponding T 2 statistics are given in Table 3.1. We plot the six parameter estimates

    for each of the twenty-four boards in Figure 3.3. The plots of a1 and b1 expose a

    potential outlier in board 15. The plots of a2 and b2 reveal potential outliers in

    boards 4, 18, and 24. Boards 4 and 15 also appear as potential outliers in the plot of

    d.

    26

  • Table 3.1: Estimated Parameter Values and T 2 Statistics for the VDP Data

    Board â1 â2 b̂1 b̂2 ĉ d̂ T2P;i T

    2D;i T

    2IPP;i T

    2MVE;i

    1 6560 3259 5.63 4.40 45.98 0.29 2.65 1.91 2261.01 6.002 470 291 3.01 2.74 42.08 0.32 7.56 5.27 4310.54 6.973 1812 2871 3.99 5.02 47.66 0.34 5.83 7.17 11255.64 8.644 6171 15009 4.25 7.39 46.63 0.39 12.21 17.28 4084.67 1131.815 4963 2251 5.14 4.20 43.43 0.30 1.65 2.27 1031.75 2.886 4556 3758 5.28 4.72 40.13 0.30 8.49 13.03 20105.39 9.837 5542 3815 5.25 5.00 44.15 0.31 2.15 3.49 211.32 3.588 3664 2979 4.89 4.41 44.06 0.30 0.79 0.97 129.64 2.699 28041 8872 7.58 4.95 43.22 0.26 4.62 7.10 3137.89 385.0310 1640 1207 4.17 3.39 41.84 0.28 4.30 5.05 5882.99 4.6111 3492 1031 5.82 3.17 46.06 0.25 8.66 8.95 2964.00 10.0012 915 750 3.45 3.52 44.37 0.32 1.80 1.99 334.45 2.2213 989 1392 3.58 4.05 45.47 0.32 3.42 4.42 1830.18 5.1814 1474 620 4.82 3.29 42.52 0.27 3.28 4.50 3218.19 7.0415 129068 5420 12.40 3.33 45.90 0.15 21.45 22.18 2048.70 17018.9116 10166 3822 5.83 4.86 44.19 0.30 3.83 5.60 266.80 12.9317 1483 603 4.07 3.26 44.83 0.30 2.30 2.53 663.30 2.3618 31156 31069 7.70 5.94 46.46 0.27 14.55 19.75 3113.48 8221.0019 418 198 3.22 2.67 42.84 0.30 4.58 3.90 1915.40 5.1620 3207 4741 4.88 5.02 44.45 0.30 5.34 5.59 23.82 34.0021 672 773 3.37 3.37 44.46 0.31 2.64 3.42 471.33 2.7922 3520 1807 5.10 4.01 45.52 0.29 1.71 1.37 1324.44 1.7323 1979 845 4.24 3.66 45.53 0.32 4.45 4.85 1843.91 7.3824 6095 26778 5.41 6.67 44.46 0.31 9.75 10.55 416.01 6676.21

    We simulated UCLs for all four of the T 2 statistics to achieve an overall probability

    of a signal equal to 0.05 for m = 24 boards. In our simulations, we sampled from

    a multivariate normal distribution of dimension six, mean vector zero, and variance-

    covariance matrix I, since the in-control performance of the methods does not depend

    on the assumed in-control parameter vector or the variance-covariance matrix. We

    repeated our simulation 200,000 times for each T 2 statistic, giving a standard error

    for the estimated control limits less than 0.0005. The four UCL values are 14.72,

    23.33, 28.22, and 65.37, for the T 2P , T2D, T

    2IPP , and T

    2MV E control charts, respectively.

    27

  • Figure 3.3: Nonlinear Regression Parameter Estimates a1, a2, b1, b2, c, and d byBoard for the VDP Data

    5 10 15 20

    040

    000

    1000

    00

    Board

    a 1

    5 10 15 20

    010

    000

    2500

    0

    Board

    a 2

    5 10 15 20

    46

    810

    12

    Board

    b 1

    5 10 15 20

    34

    56

    7

    Board

    b 2

    5 10 15 20

    4042

    4446

    Board

    c

    5 10 15 20

    0.15

    0.25

    0.35

    Board

    d

    In Section 3.1.3 we gave theoretical UCLs for the T 2P , T2D, and T

    2IPP control charts.

    For m = 24 boards and p = 6 parameters, the theoretical UCLs are 14.71, 11.85,

    and 20.63 for the T 2P , T2D, and T

    2IPP control charts, respectively. The exact marginal

    distribution of the T 2P,i statistic is known, thus the theoretical and simulated UCL

    values are very similar. On the other hand, the exact marginal distribution of the

    T 2D,i statistic is not known, so we used instead the correction given by Mason and

    Young (2002, pp. 26-27). The large difference between the simulated and theoretical

    UCL for the T 2D control chart shows that the approximate marginal distribution is

    28

  • inadequate. The UCL of the T 2IPP control chart was computed based on the marginal

    asymptotic distribution of the T 2IPP,i statistic because the small-sample distribution

    is unknown. For our VDP example, the number of samples in the baseline dataset

    is only m = 24 boards. The theoretical UCL values become more exact as m gets

    larger.

    In Phase I analysis, we are interested in identifying “outlying” or out-of-control

    boards or a shift in the process which might affect the estimation of in-control pa-

    rameters. We compared the four T 2 control charts for assessing process stability and

    identifying outlying profiles. In Figure 3.4 we illustrate all four T 2 control charts for

    the VDP data.

    Figure 3.4: The T 2 Control Charts for the VDP Data. (a) The T 2P control chart basedon the pooled sample covariance matrix, (b) T 2D control chart based on the successivedifferences estimator, (c) T 2IPP control chart based on the intra-profile pooling method,and (d) T 2MV E control chart based on the minimum volume ellipsoid, with UCL valuesof 14.72, 23.33, 28.22, and 65.37, respectively.

    5 10 15 20

    05

    1015

    2025

    (a)

    Board

    5 10 15 20

    05

    1015

    2025

    (b)

    Board

    5 10 15 20

    050

    0015

    000

    (c)

    Board

    5 10 15 20

    050

    0010

    000

    (d)

    Board

    29

  • The T 2P control chart based on the pooled sample variance-covariance matrix esti-

    mator indicates that board 15 has the only out-of-control profile, although the profile

    for board 18 is borderline. The T 2D chart based on the successive differences estimator

    does not produce an out-of-control signal. Note that the T 2D,i statistic accentuates

    the same outlying observations of the T 2P chart, but has a larger UCL. As discussed in

    Sullivan and Woodall (1996), the T 2P control chart has greater power to detect isolated

    outlying observations than the T 2D control chart based on the successive differences

    variance-covariance matrix estimator, however the T 2D chart is better for detecting a

    sustained shift in the mean vector. For this dataset, there is no apparent sustained

    shift in the regression parameter vector.

    In the T 2IPP control chart based on the intra-profile pooling variance-covariance

    matrix estimator, all the T 2IPP,i statistics are above the UCL except for board 20,

    indicating that most of the profiles are significantly different from each other in the

    statistical sense. Recall that SIPP in the T2IPP,i statistic is the average of the m

    within-profile variance-covariance matrices of the β̂i vectors. For this dataset, the

    within-profile variability is much smaller than the between-profile variability, causing

    the T 2IPP,i statistics to be very large. The use of this method is appropriate only if

    there is no expected common cause variation between profiles. We expect that in

    most applications there will be some common cause variation between profiles.

    The T 2MV E control chart based on the MVE estimator indicates that boards 4,

    9, 15, 18, and 24 have outlying profiles. The most pronounced outlier is board 15,

    which the T 2P chart also indicated as the most severe outlier. As shown by Vargas

    (2003), the T 2MV E control chart is very powerful in detecting multivariate outliers.

    Investigating the table of parameter estimates for these boards, given in Table 3.1,

    it seems reasonable that the boards 15 and 18 are outliers, with boards 4, 9, and 24

    worthy of further investigation.

    30

  • As discussed in Section 3.1.5, an alternative approach to modeling the profiles

    with a parametric curve is to employ nonparametric smoothing techniques to model

    the profiles. Walker and Wright (2002) employed spline smoothing with 16 degrees of

    freedom to model the twenty-four boards of the VDP data. We replicated their spline

    fits to each profile. After obtaining the spline fits to each profile, ẏij, i = 1, . . . , m; j =

    1, . . . , n, the average spline, ỹj, is calculated. For example, the spline fit to board 1

    and the average spline are illustrated in Figure 3.5.

    Figure 3.5: Spline Fit of Board 1 (Above) and Average Spline (Below) for the VDPData

    0.0 0.1 0.2 0.3 0.4 0.5 0.6

    4045

    5055

    6065

    Depth (inches)

    Den

    sity

    (lbs

    ft3)

    The spline fit with 16 degrees of freedom provides a concise summary of the shape

    of the profile from board 1. The average spline fit is systematically lower than the

    spline fit to board 1. In order to determine which boards are in-control we calculated

    31

  • dissimilarity metrics as given in Section 3.1.5. Since the metrics Mi2 and Mi3 differ

    only by a constant, it is not helpful to consider both metrics simultaneously. Instead

    we calculate the metrics Mi1, Mi3, Mi4, and Mi5, and then employ an I-chart based

    on the moving range to establish control limits, as suggested by researchers at Boeing

    (1998). We plot each metric versus i with associated control limits to obtain control

    charts. The four charts are given in Figure 3.6.

    Figure 3.6: Control Charts on Metrics: (a) Mi1, the Maximum Deviation, (b) Mi3,the Maximum Absolute Deviation, (c) Mi4, the Sum of Squared Differences, and (d)Mi5, the Mean Absolute Deviation for the VDP Data

    5 10 15 20

    −10

    −50

    510

    (a)

    Board

    Max

    imum

    Dev

    iatio

    n

    5 10 15 20

    −50

    510

    (b)

    Board

    Max

    imum

    Abs

    olut

    e De

    viatio

    n

    5 10 15 20

    −200

    00

    2000

    6000

    (c)

    Board

    Sum

    of S

    quar

    ed D

    iffer

    ence

    s

    5 10 15 20

    −20

    24

    (d)

    Board

    Mea

    n Ab

    solu

    te D

    evia

    tion

    The charts based on metrics Mi1 and Mi4 both give the same conclusion, that all

    the profiles of the boards are in-control. This is not surprising since Mi4 is the absolute

    32

  • value of Mi1, but both are given for illustrative purposes. The most extreme value

    of the metrics came from board 14, with values of M14,1 = −5.79 and M14,4 = 5.79 .This value represents the maximum (absolute) deviation of the spline fit to board 14

    from the average spline fit.

    Similarly, the charts based on metrics Mi3 and Mi5 both give the same conclusion,

    that the profile for board 6 is out-of-control. Referring to Figure 3.1, board 6 is

    the one with the profile that is consistently lower than all other boards. The next

    most extreme value of the two metrics is that of board 3, although it does not give

    an out-of-control signal. Again, referring to Figure 3.1, board 3 is the one with the

    profile that is consistently higher than all the other boards. It is apparent that these

    two metrics measure how consistently different each profile is from the average profile

    across the depth values, whereas metrics Mi1 and Mi4 measure the greatest extent to

    which a profile is from the average at any particular depth value. It is important to

    note that the results for the control charts on the metrics (Figure 3.6) do not show

    the same results as the control charts based on the regression estimators in Figure

    3.4. If the profile can be adequately represented by a parametric model, then this, in

    general, will lead to more effective charts.

    In addition to monitoring the regression parameter vectors of the profiles in a

    Phase I analysis, we should monitor the variation about the profiles to check for

    stability. As mentioned in Section 2.4, we recommend using the methods of Wludyka

    and Nelson (1997) to monitor the variance σ2. Use of their method is appropriate

    when the error terms within a profile are independent. In our VDP example, however,

    the within-profile density measurements are spatially correlated. A more appropriate

    control chart in this case to monitor the process variance σ2 is a topic for further

    research.

    33

  • 3.3 Autocorrelation

    Engineering applications that give rise to nonlinear profile data may lead to autocor-

    related error terms. A common source of autocorrelated errors is the spatial or serial

    manner in which data are collected. The VDP data, for example, is spatially corre-

    lated because the density measurements are taken at close intervals along the vertical

    depth of the particleboard. On the other hand, some nonlinear profiles may have

    independent error terms. One example of this is typical dose-response data where

    several doses of a particular drug are administered to different subjects and their

    responses are measured. The subsequent error terms in the nonlinear dose-response

    curve are typically assumed to be independent.

    When the error terms are autocorrelated due to either serial, spatial or any other

    effects, the correlation structure should be taken into account in the analysis. Failure

    to do so might yield misleading results in some cases, particularly with the control

    chart to monitor σ2. In the example of Section 3.2, we estimated parameters of a

    nonlinear regression model for each board. For our nonlinear model we assumed that

    the errors ²ij are i.i.d. For the VDP data, it may be reasonable to assume that the

    ²ij are correlated. If this is the case, perhaps an alternative approach would be to

    employ either nonlinear mixed model methods or generalized estimating equations

    (GEE) methodology. Both methods can be used to estimate the mean function, or

    profile, while accounting for autocorrelation in the error structure. A more detailed

    treatment of these methods can be found in Schabenberger and Pierce (2002) and

    Hardin and Hilbe (2003). In the context of analyzing nonlinear profiles for Phase I

    applications, this approach is a topic that requires further investigation.

    34

  • 3.4 Discussion

    In Phase I, we are interested in identifying outlying observations as well as identifying

    step or ramp shifts in the mean vector over time. As shown by Vargas (2003), the

    robust variance-covariance matrix and mean vector estimators employed in the T 2MV E,i

    statistic are very powerful in detecting multivariate outliers, but are not powerful in

    detecting a step shift. However, the opposite is true of the T 2D,i statistic. As shown

    by Sullivan and Woodall (1996), the T 2D chart is powerful in detecting a step shift,

    but not powerful in detecting multivariate outliers. One possible alternative is to

    employ both the T 2D and T2MV E charts simultaneously, the former chart sensitive to

    step shifts and the latter sensitive to outliers. However, in examining both charts

    simultaneously, one must be cautious of inflating the false alarm probability. This

    approach is also a topic for research, as discussed in Chapter 6.

    We have not given a detailed treatment of the nonparametric approaches to moni-

    toring profiles discussed in Section 3.1.5. Rather, we have only described some meth-

    ods that have been proposed and then illustrated their use with the VDP data. Some

    issues that need to be addressed, for example, are the best nonparametric estimation

    technique for a given scenario, the best metrics to apply, the strengths and weaknesses

    of each metric, and the distributional properties of the metrics in order to establish

    valid control limits. It is our hope that the present work will generate interest in

    investigating these and other unresolved issues.

    The field of profile monitoring using control charts has potential to extend statis-

    tical process control to a wide variety of engineering and pharmaceutical applications.

    With the increasing ease and efficiency in which processes and products can be mea-

    sured, there is a need for statistical methodology to be developed which can accom-

    modate the growing needs of industry. We have encountered a number of engineering

    applications in which a response curve is needed to assess quality. In some cases, the

    35

  • shape of the response curve can be well-represented by a parametric nonlinear regres-

    sion function. In this paper we have developed control chart methodology to monitor

    such nonlinear profiles for Phase I applications. When a profile cannot be easily de-

    scribed by a parametric function, nonparametric methods may be applied. Our VDP

    example shows, however, that the parametric and nonparametric approaches do not

    always lead to the same conclusions regarding outlying profiles.

    36

  • Chapter 4

    Distribution of the T 2 StatisticBased on the SuccessiveDifferences Covariance MatrixEstimator

    4.1 The T 2 Statistic

    In a Phase I analysis, we begin with an HDS consisting of m independent vectors

    of dimension p observed over time, where p is the number of quality characteristics

    that are being measured, and p < m. We make a standard assumption that when

    the process is in-control the observation vectors, xi, i = 1, . . . , m, are independent

    and identically distributed multivariate normal random vectors with common mean

    vector and covariance matrix, i.e.,

    xi ∼ Np(µ,Σ).

    For example, in the context of nonlinear profile monitoring discussed in Chapter 3,

    each xi is equal to β̂i, i = 1, . . . , m. It is useful to define the m× p HDS matrix X as

    X =

    x′1x′2...

    x′m

    .

    37

  • The Hotelling’s T 2 statistic measures the Mahalanobis distance of the correspond-

    ing vector from the sample mean vector. The general form of the statistic is

    T 2i = (xi − x̄)′ S−1 (xi − x̄) ,

    where x̄ = 1m

    ∑mi=1 xi and S is some estimator of Σ.

    A common choice for S is the sample variance-covariance estimator given by

    SP =1

    m− 1m∑

    i=1

    (xi − x̄) (xi − x̄)′ .

    The T 2 statistics based on SP are then

    T 2P,i = (xi − x̄)′ S−1P (xi − x̄) , i = 1, 2, . . . , m (4.1)

    As shown in Chapter 3, the exact distribution of T 2P,i is proportional to a beta distri-

    bution, i.e.,

    T 2P,im

    (m− 1)2 ∼ BETA(

    p

    2,m− p− 1

    2

    ); i = 1, . . . , m. (4.2)

    An alternative choice of S is one based on SD of Equation (3.8), where β̂i is

    replaced with xi. The resulting T2 statistics based on SD are given by

    T 2D,i = (xi − x̄)′ S−1D (xi − x̄) . (4.3)

    As noted in Sullivan and Woodall (1996), Holmes and Mergen (1993) incorrectly

    specify the Phase I UCL of a T 2 chart based on T 2D,i statistics by applying con-

    trol limits based on a Phase II analysis. Sullivan and Woodall (1996) proposed an

    approximate distribution for T 2D,i as

    T 2D,im

    (m− 1)2 ∼ BETA(

    p

    2,f − p− 1

    2

    ), (4.4)

    where f = 2(m−1)2

    3m−4 . Mason and Young (2002, pp. 26-27) suggested an adjustment to

    this approximation, replacing each m in Equation (4.4) with