review and recommendation of methods for …frey/reports/mokhtari_frey_2005b.pdfi preface this is...

179
Review and Recommendation of Methods for Sensitivity and Uncertainty Analysis for the Stochastic Human Exposure and Dose Simulation (SHEDS) Models Volume 2: Evaluation and Recommendation of Methodology for Conducting Sensitivity Analysis in Probabilistic Models 2004-206-01 Prepared by: Amirhossein Mokhtari H. Christopher Frey Department of Civil, Construction, and Environmental Engineering North Carolina State University Raleigh, NC Prepared for: Alion Science and Technology 1000 Park Forty Plaza Durham, NC June 30, 2005

Upload: doannhu

Post on 04-May-2018

215 views

Category:

Documents


0 download

TRANSCRIPT

Review and Recommendation of Methods for Sensitivity and Uncertainty Analysis for the Stochastic Human

Exposure and Dose Simulation (SHEDS) Models

Volume 2: Evaluation and Recommendation of Methodology for Conducting

Sensitivity Analysis in Probabilistic Models

2004-206-01

Prepared by:

Amirhossein Mokhtari H. Christopher Frey

Department of Civil, Construction, and Environmental Engineering North Carolina State University

Raleigh, NC

Prepared for:

Alion Science and Technology 1000 Park Forty Plaza

Durham, NC

June 30, 2005

i

Preface

This is one of a two volume series of reports on the topic of review and recommendation

of methods for sensitivity and uncertainty analysis for the Stochastic Human Exposure and Dose

Simulation (SHEDS) models.

The first volume provides a comprehensive review of methods for sensitivity and

uncertainty analysis, with a focus on methods that are relevant to probabilistic models based

upon Monte Carlo simulation or similar techniques for propagation of distributions for variability

and uncertainty in model inputs in order to estimate variability and uncertainty in model outputs.

The methods included in the review are those that are considered to be “available,” which is

interpreted to be methods of practical significance as opposed to all possible methods that have

been proposed but not tested in practice. For each method, there is a description of the method,

followed by a discussion of the advantages and disadvantages of the method. A framework for

selection of sensitivity analysis methods is presented, leading to recommendations for a more

narrow set of such methods that merit more detailed evaluation.

The second volume proposes and applies a methodology for evaluation of the selected

sensitivity analysis methods based upon application of each method to a modeling testbed. The

testbed is a simplified version of a typical SHEDS model. A case study scenario was defined

that includes multiple time scales (e.g., daily, monthly). Seven sensitivity analysis methods were

applied to the testbed, including Pearson correlation, Spearman correlation, sample regression,

rank regression, Analysis of Variance (ANOVA), Fourier Amplitude Sensitivity Test (FAST),

and Sobol’s method. The sensitivity analysis results obtained from these seven methods were

compared. On the basis of these quantitative results, recommendations were made for methods

that offer promise for application to SHEDS models. The statistically-based methods are often

ii

readily available features of commonly available software packages. However, FAST and

Sobol’s method are less readily available. Therefore, algorithms are presented for these two

methods. Furthermore, recommendations are made for additional research and development of

sensitivity analysis methods for application to the SHEDS models.

iii

TABLE OF CONTENTS

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

2. OVERVIEW OF THE STOCHASTIC HUMAN EXPOSURE AND DOSE SIMULATION (SHEDS) MODELS ................................................................................5 2.1 Overview of the SHEDS-Pesticides Model .............................................................5

2.2 Main Characteristics of the SHEDS-Pesticide Model .............................................8

2.3 Simplified Version of the SHEDS-Pesticides Model ..............................................9

3. OVERVIEW OF UNCERTAINTY AND SENSITIVITY ANALYSIS METHODS .......................................................................................................................13 3.1 Methods for Propagation and Quantification of Variability and

Uncertainty.............................................................................................................13 3.1.1 Analytical Propagation Techniques...........................................................13 3.1.2 Approximation Methods.............................................................................14 3.1.3 Numerical Propagation Techniques ..........................................................14

3.2 Sensitivity Analysis ...............................................................................................16 3.2.1 Correlation Analysis ..................................................................................16 3.2.2 Linear Regression Analysis........................................................................17 3.2.3 Analysis of Variance ..................................................................................17 3.2.4 Sobol’s Method ..........................................................................................18 3.2.5 Fourier Amplitude Sensitivity Test (FAST)................................................20

4. DEFINING CASE SCENARIOS FOR SENSITIVITY ANALYSIS ..........................23 4.1 Probabilistic Dimensions .......................................................................................24

4.2 Susceptible Subpopulation.....................................................................................25

4.3 Identification of Pathway of Interest......................................................................26

4.4 Time Scales of the Model Simulation....................................................................26

5. RESULTS OF SENSITIVITY ANALYSIS FOR CASE STUDIES ...........................31 5.1 Model Application .................................................................................................31

5.2 Correlation Analysis ..............................................................................................33 5.2.1 Correlation Coefficients Results ................................................................34 5.2.2 Rankings for Selected Inputs......................................................................47 5.2.3 Comparison of Mean Ranks.......................................................................52 5.2.4 Summary of Results Based on Correlation Coefficients ............................55

5.3 Regression Analysis...............................................................................................58 5.3.1 Coefficient of Determination......................................................................58 5.3.2 Regression Coefficients..............................................................................61 5.3.3 Rankings for Selected Inputs......................................................................67

iv

5.2.4 Comparison of Mean Ranks.......................................................................71 5.2.5 Summary of Results Based on Regression-Based Methods .......................74

5.4 Analysis of Variance..............................................................................................76 5.4.1 F Values .....................................................................................................77 5.4.2 Rankings for Selected Inputs......................................................................78 5.4.3 Comparison of Mean Ranks.......................................................................80 5.4.5 Summary of Results Based on ANOVA ......................................................82

5.5 Sobol’s Method......................................................................................................83 5.5.1 General Insight Regarding Contribution of Inputs to the Output

Variance.....................................................................................................83 5.5.2 Sensitivity Indices.......................................................................................89 5.5.3 Ranking for Selected Inputs .......................................................................93 5.5.4 Comparison of Mean Ranks.......................................................................95 5.5.5 Summary of Results Based on Sobol’s Method..........................................98

5.6 Fourier Amplitude Sensitivity Test........................................................................99 5.6.1 General Insight Regarding Contribution of Inputs to the Output

Variance.....................................................................................................99 5.6.2 Sensitivity Indices.....................................................................................101 5.6.3 Ranking for Selected Inputs .....................................................................105 5.6.4 Comparison of Mean Ranks.....................................................................106 5.6.5 Summary of Results Based on FAST........................................................109

5.7 Comparison of Results for Selected Sensitivity Analysis Methods.....................110

6. CONCLUSIONS AND RECOMMENDATIONS.......................................................115 6.1 Main Characteristics of the SHEDS Models Relevant to the Process of

Choosing Appropriate Sensitivity Analysis Methods..........................................115

6.2 Available Sensitivity Analysis Methods..............................................................115

6.3 Sensitivity Analysis Methods for Application to the SHEDS Models ................116

6.4 Recommendations................................................................................................125

REFERENCES...........................................................................................................................131

APPENDIX A .............................................................................................................................139

APPENDIX B .............................................................................................................................149

v

LIST OF FIGURES

Figure 2-1. Schematic Diagram of the SHEDS-Pesticides Simulation Process. ...................7 Figure 2-2. Schematic Diagram for the Simplified SHEDS-Pesticides Model. ....................9 Figure 4-1. Components of a Scenario for Sensitivity Analysis of the Simplified

SHEDS-Pesticides Model. ............................................................................24 Figure 4-2. Probability Distribution of Different Average Exposure Pathways in 30

Days of Model Simulation. ...........................................................................27 Figure 4-3. Example of Probabilistic Results for Variation in the Total Exposure

for Selected Random Individuals: (a) Scenario I-Daily Exposure; (b) Scenario II-Incremental Change in Daily Exposure; and (c) Scenario III-Cumulative Exposure. .............................................................................29

Figure 5-1. Schematic Algorithm for Application of Selected Sensitivity Analysis Methods to the Simplified SHEDS-Pesticides Model. .................................33

Figure 5-2. Absolute Values and Corresponding 95% Confidence Intervals of Pearson and Spearman Correlation Coefficients for Re sidue Decay Rate as an Input with Monthly Sampling Frequency for: (a) Scenario I: Daily Exposure; (b) Scenario II: Incremental Change in Daily Exposure; and (c) Scenario III: Cumulative Daily Exposure. ............35

Figure 5-3. Absolute Values and Corresponding 95% Confidence Intervals of Pearson and Spearman Correlation Coefficients for Probability of Washing Body as an Input with Monthly Sampling Frequency for: (a) Scenario I: Daily Exposure; (b) Scenario II: Incremental Change in Daily Exposure; and (c) Scenario III: Cumulative Daily Exposure. ........41

Figure 5-4. Absolute Values and Corresponding 95% Confidence Intervals of Pearson and Spearman Correlation Coefficients for Fraction of Chemicals Available for Transfer as an Input with a Daily Sampling Strategy for: (a) Scenario I: Daily Exposure; (b) Scenario II: Incremental Change in Daily Exposure; and (c) Scenario III: Cumulative Daily Exposure..........................................................................44

Figure 5-5. Variation in Rank of Residue Decay Rate as an Input with Monthly Sampling Strategy with Respect to Time Based on: (a) Pearson; and (b) Spearman Correlation Coefficients. ........................................................48

Figure 5-6. Variation in Rank of Fraction of Chemical Available for Transfer as an Input with Daily Sampling Strategy with Respect to Time Based on: (a) Pearson; and (b) Spearman Correlation Coefficients. .............................50

Figure 5-7. Comparison of Mean Ranks and Range of Ranks Representing Minimum and Maximum Ranks of Inputs in 30 Days of Model Simulation for the Three Temporal Scenarios Based on: (a) Pearson Correlation Coefficients; and (b) Spearman Correlation Coefficients. ........53

Figure 5-8. Variation in Coefficient of Determination, R2, with Respect to Time based on Sample and Rank Regression Analyses for: (a) Scenario I: Daily Exposure; (b) Scenario II: Incremental Change in Daily Exposure; and (c) Scenario III: Cumulative Exposure. ................................60

Figure 5-9. Absolute Values and Corresponding 95% Confidence Intervals of Sample and Rank Regression Coefficients for Residue Decay Rate as

vi

an Input with a Monthly Sampling Strategy for: (a) Scenario I: Daily Exposure; (b) Scenario II: Incremental Change in Daily Exposure; and (c) Scenario III: Cumulative Daily Exposure. ......................63

Figure 5-10. Absolute Values and Corresponding 95% Confidence Intervals of Sample and Rank Regression Coefficients for Probability of Washing Body as an Input with a Monthly Sampling Strategy for: (a) Scenario I: Daily Exposure; (b) Scenario II: Incremental Change in Daily Exposure; and (c) Scenario III: Cumulative Daily Exposure. ........66

Figure 5-11. Absolute Values and Corresponding 95% Confidence Intervals of Sample and Rank Regression Coefficients for Fraction of Chemicals Available for Transfer as an Input with a Daily Sampling Strategy for: (a) Scenario I: Daily Exposure; (b) Scenario II: Incremental Change in Daily Exposure; and (c) Scenario III: Cumulative Daily Exposure. ......................................................................................................68

Figure 5-12. Variation in Rank of Residue Decay Rate as an Input with Monthly Sampling Strategy with Respect to Time Based on: (a) Sample Regression Analysis; and (b) Rank Regression Analysis. ............................69

Figure 5-13. Variation in Rank of Fraction of Chemical Available for Transfer as an Input with a Daily Sampling Strategy with Respect to Time Based on: (a) Sample Regression Analysis; and (b) Rank Regression Analysis.........................................................................................................70

Figure 5-14. Comparison of Mean Ranks and Range of Ranks Representing Minimum and Maximum Ranks of Inputs in 30 Days of Model Simulation for the Three Temporal Scenarios Based on: (a) Sample Regression Analysis; and (b) Rank Regression Analysis. ............................72

Figure 5-15. Variation of F Values with Respect to Time for the Three Temporal Scenarios for: (a) Residue Decay Rate; and (b) Fraction of Chemicals Available for Transfer. ................................................................78

Figure 5-16. Variation of Ranks with Respect to Time for the Three Temporal Scenarios for: (a) Residue Decay Rate; and (b) Fraction of Chemicals Available for Transfer. ................................................................79

Figure 5-17. Comparison of Mean Ranks and Range of Ranks of Inputs in 30 Days for the Three Temporal Scenarios Based on ANOVA. ................................81

Figure 5-18. Variation in Average Sum of Main Effects of Inputs in 500 Bootstrap Simulations and Corresponding Uncertainty Ranges with Respect to Time for the Three Temporal Scenarios. (Intervals shown represent plus or minus one standard error of the mean value)....................................85

Figure 5-19. Contribution Selected Inputs to the Output Variance in Scenario I for Selected Days: (a) Day 1; (b) Day 15; and (c) Day 30. ...............................87

Figure 5-20. Average Main and Total Effects and Corresponding Uncertainty Ranges Based on 500 Bootstrap Simulations for Residue Decay Rate for: (a) Scenario I-Daily Exposure; (b) Scenario II-Incremental Change in Daily Exposure; and (c) Scenario III-Cumulative Exposure. ......................................................................................................91

Figure 5-21. Average Main and Total Effects and Corresponding Uncertainty Ranges Based on 500 Bootstrap Simulations for Residue Decay Rate

vii

for: (a) Scenario I-Daily Exposure; (b) Scenario II-Incremental Change in Daily Exposure; and (c) Scenario III-Cumulative Exposure. ......................................................................................................92

. 92 Figure 5-22. Variation of Ranks with Respect to Time for the Three Temporal

Scenarios for: (a) Residue Decay Rate; and (b) Fraction of Chemicals Available for Transfer Based on the Sobol’s Method.................94

Figure 5-23. Comparison of Mean Ranks and Range of Ranks Based on the Sobol’s Method for Inputs in 30 Days for the Three Temporal Scenarios Based on: (a) Main Effects; and (b) Total Effects. ......................................97

Figure 5-24. Variation in Sum of Main Effects of Inputs Based on FAST with Respect to Time for: (a) Daily; (b) Incremental; and (c) Cumulative Total Exposure Scenarios. ..........................................................................100

Figure 5-25. Main and Total Effects for Residue Decay Rate Based on FAST for: (a) Scenario I-Daily Exposure; (b) Scenario II-Incremental Change in Daily Exposure; and (c) Scenario III-Cumulative Exposure. .................103

Figure 5-26. Main and Total Effects for Fraction of Chemicals Available for Transfer Based on FAST for: (a) Scenario I-Daily Exposure; (b) Scenario II-Incremental Change in Daily Exposure; and (c) Scenario III-Cumulative Exposure. ...........................................................................104

Figure 5-27. Variation of Ranks with Respect to Time for the Three Temporal Scenarios for: (a) Residue Decay Rate; and (b) Fraction of Chemicals Available for Transfer Based on FAST. ...................................106

Figure 5-28. Comparison of Mean Ranks and Range of Ranks Based on the FAST for Inputs in 30 Days for the Three Temporal Scenarios Based on: (a) Main Effects; and (b) Total Effects. ......................................................108

Figure 5-29. Comparison of the Mean Ranks and Corresponding Range of Ranks in 30 Days of Model Simulation Based on Pearson Correlation Analysis (PCA), Spearman Correlation Analysis (SCA), Sample Regression Analysis (SRA), Rank Regression Analysis (RRA), Analysis of Variance (ANOVA), Sobol’s Method, and FAST for: (a) Inputs with Monthly Sampling Frequencies; and (b) Inputs with Daily Sampling Frequencies in Scenario II ................................................112

Figure A.1. Schematic Diagram of the Simplified SHEDS Model. ..................................140 Figure B-1: Matlab Code for Assigning Frequencies to Daily Inputs. ..............................155 Figure B-2: Matlab Code for Selecting a Search Curve for Inputs with Uniform,

Loguniform, and Normal Distributions. .....................................................156 Figure B-3: Matlab Code for Selecting a Search Curve for Inputs with Lognormal

Distribution. ................................................................................................157 Figure B-4: Matlab Code for Estimating the Output Variance..........................................158 Figure B-5: Matlab Code for Estimating the Fourier Coefficients. ...................................159 Figure B-6: Matlab Code for Assigning New Frequencies to Daily Inputs Used in

Estimation of Total Sensitivity Indices.......................................................160 Figure B-7. Flow Diagram for Application of the Sobol’s Method. .................................166 Figure B-8. Forming Mj and Nj Matrices for Input xj with Monthly Sampling

Frequency....................................................................................................167

viii

Figure B-9. Forming Mj and Nj Matrices for Input xj with Daily Sampling Frequency....................................................................................................168

Figure B-10. Matlab Code for Forming M and N Matrices for Inputs with Monthly and Daily Sampling Frequencies. ...............................................................169

ix

LIST OF TABLES

Table 2-1. Examples for Input Assumptions for the Simplified SHEDS-Pesticides Model ........11 Table 4-1. Example of Probabilistic Results for Daily, Incremental Change in Daily, and

Cumulative Total Exposure for the Random Individual ‘a’ shown in Figure 4-3 ..........................................................................................................................30

Table 6-1. Comparison of Various Sensitivity Analysis Methods .............................................117 Table A.1. Input Assumptions for the Simplified SHEDS model ..............................................146

1

1. INTRODUCTION

The purpose of this project is to identify, evaluate, and recommend methods for

sensitivity analysis applicable to the Stochastic Human Exposure and Dose Simulation (SHEDS)

risk assessment models. The EPA SHEDS models are aggregate, probabilistic and physically-

based human exposure models that simulate variability and uncertainty in cumulative human

exposure and dose (EPA, 2000; Price et al., 2003). Understanding key sources of variability can

guide the identification of significant subpopulations that merit more focused study and in

developing approaches for risk management. In contrast, knowing the key sources of uncertainty

can aid in determining whether additional research or alternative measurement techniques are

needed to reduce uncertainty (Cullen and Frey, 1999).

Volume I of this series of report provides an overview of motivations for conducting

sensitivity and uncertainty analyses. Sensitivity analysis is recognized as an essential component

of analyses for complex models (Helton, 1997). Sensitivity analysis highlights the inputs that

have the greatest influence on the results of a model; therefore, it provides useful insights for

model builders and users (McCarthy et al., 1995). Sensitivity analysis provides insight regarding

which model input contributes the most to uncertainty, variability, or both, for a particular model

output. Insights from sensitivity analysis can be used for: (1) identification of key sources of

uncertainty; (2) identification of key controllable sources of variability; and (3) model

refinement, verification, and validation (Mokhtari and Frey, 2005).

Sensitivity analyses of risk models are used to identify inputs that matter the most with

respect to exposure or risk and aid in developing priorities for risk mitigation and management

(Baker et al., 1999; Jones, 2000). Sensitivity analysis can identify important uncertainties.

Additional data collection or research can be prioritized based on the key uncertainties in order

2

to reduce the model output uncertainty (Cullen and Frey, 1999). Knowledge of key controllable

sources of variability is useful in identifying control measures to reduce exposure or risk.

Sensitivity analysis has been used for verification and validation purposes during the

process of model development and refinement (e.g., Kleijnen, 1995; Kleijnen and Sargent, 2000;

Fraedrich and Goldberg, 2000). Sensitivity analysis can be used for verification by assessing

whether the model output responds appropriately to a change in model inputs. Sensitivity

analysis can validate a model by determining the degree to which a model is an accurate

representation of the real world. Sensitivity analysis also can be used to evaluate the robustness

of model results (e.g., Philips et al., 2000; Ward and Carpenter, 1996; Limat et al., 2000;

Manheim, 1998; Saltelli et al., 2000).

There are many sensitivity analysis methods applied in various scientific fields, including

engineering, economics, physics, social sciences, medical decision making, and others (e.g.,

Baniotopoulos, 1991; Cheng, 1991; Merz et al., 1992; Helton and Breeding, 1993; Beck et al.,

1997; Agro et al., 1997; Kewley et al., 2000; Oh and Yang, 2000). Given the myriad of

sensitivity analysis methods, there is a need for insight regarding which methods to choose and

regarding how to apply preferred methods. In Volume 1, approximately a dozen available

sensitivity analysis methods were identified and reviewed, including their advantages and

disadvantages. Based upon the review, a selected set of seven sensitivity analysis methods was

selected for more detailed and quantitative evaluation. These methods include Pearson and

Spearman correlation , sample and rank regression, analysis of variance (ANOVA), Fourier

amplitude sensitivity test (FAST), and Sobol’s method. In this volume, these selected methods

are applied to a testbed in order to facilitate comparisons based upon quantitative results. The

testbed is a simplified version of a typical SHEDS model.

3

A key objective of the sensitivity analysis of the SHEDS models was to quantify the

contribution of individual inputs to the variance in the selected model output. Thus, sensitivity

analysis methods were selected for further evaluation based on their capability in proving

insights with respect to the key objective. Among the selected techniques, FAST and Sobol’s

method are variance-based methods and provide quantitative measures for contributions of

inputs to the output variance. Although selected sampling-based methods do not provide

measures regarding the contributions of inputs to the output variance, comparison of the results

based on these methods and the two variance-based techniques can provide insights regarding

trade-off between application of commonly used sensitivity analysis methods and more

sophisticated techniques.

The key questions that are addressed in this report include the following:

(1) What are the main characteristics of the SHEDS models that are relevant to the

process of choosing appropriate sensitivity analysis methods?

(2) Which sensitivity analysis methods described in Volume 1 are most promising

and merit more detailed evaluation?

(3) What are the recommended sensitivity analysis methods for application to the

SHEDS models?

(4) What are the key recommendations for further research based on the insights from

the analyses?

Chapter 2 briefly explains the SHEDS models including their key characteristics. Furthermore, a

simplified version of the SHEDS-Pesticides model developed for evaluation of sensitivity

analysis methods is discussed. Chapter 3 provides a brief overview of uncertainty and sensitivity

analysis methods that are of practical significance. Chapter 4 defines a set of case scenarios for

4

evaluation of selected sensitivity analysis methods. Key attributes of each scenario including

probabilistic features, susceptible subpopulations, exposure pathway of interest, and time scale of

the model simulation are explained. Chapter 5 provides results from application of selected

sensitivity analysis methods to the case scenarios defined in Chapter 4. Chapter 6 summarizes

the conclusions and recommendations based on the analyses provided in Chapter 5 and discusses

key recommendations for further research. Appendix A provides the key equations of the

simplified SHEDS-Pesticides model along with the probability distributions of model inputs.

Appendix B provides algorithms for application of two of the sensitivity analysis methods

evaluated in this report including Fourier Amplitude Sensitivity Test (FAST) and the Sobol’s

method.

5

2. OVERVIEW OF THE STOCHASTIC HUMAN EXPOSURE AND DOSE SIMULATION (SHEDS) MODELS

SHEDS is a family of models that are developed to estimate multimedia and multi-

pathway pollutant exposures of general as well as at-risk populations. The SHEDS models are

being designed to predict and diagnose complex relationships between pollutant sources and

dose received by different subpopulation (e.g., children and the elderly) (Price et al., 2003). The

SHEDS models provide estimates of variability and uncertainty in the predicted exposure

distributions and characterize factors influencing high-end exposures. These models address both

aggregate (all sources, routes, and pathways for a single chemical) and cumulative (aggregate for

multiple chemicals) exposures. The SHEDS models include SHEDS-Pesticides, SHEDS-Wood,

SHEDS-Air Toxics and SHEDS-PM (EPA, 2000). The discussion presented here is mainly based

upon SHEDS-Pesticides, because this model has typical characteristics of the SHEDS series of

models.

Sections 2.1 and 2.2 provide an overview of and briefly explain the main characteristics

of the SHEDS-Pesticides model, respectively. For the purpose of evaluation of sensitivity

analysis methods, a simplified version of the SHEDS-Pesticides model is developed. The

simplified model has the key advantage of shorter simulation time compared to the original

model. The key assumptions of the simplified SHEDS-Pesticides model are briefly explained in

Section 2.3.

2.1 Overview of the SHEDS-Pesticides Model

The EPA SHEDS-Pesticides model is an aggregate, probabilistic and physically-based

human exposure model which simulates variability and uncertainty in cumulative human

exposure and dose to pesticides. The key purposes of the model are: (1) to improve the risk

assessment process by predicting both inter-individual variability and uncertainties associated

6

with population exposure and dose distributions; (2) to improve the risk management process by

identifying critical exposure routes and pathways; and (3) to provide a framework for identifying

and prioritizing measurement needs and for formulating the most appropriate hypotheses and

designs for exposure studies. The schematic diagram of the simulation process for the SHEDS-

Pesticides model is given in Figure 2-1.

The SHEDS-Pesticides model predicts, for user-specified cohorts, exposures and doses

incurred via eating contaminated foods or drinking water, inhaling contaminated air, touching

contaminated surface residues, and ingesting residues from hand-to-mouth activities. The model

combines information on pesticide usage, human activity data, environmental residues and

concentrations, exposure and dose factors. The SHEDS-Pesticides model is limited to pesticides

and other dislodgeable compounds present on surfaces in residential environments (Price et al.,

2003).

For each individual, the SHEDS-Pesticides model can construct daily exposure and dose

time profiles for the inhalation, dietary and non-dietary ingestion, and dermal contact exposure

routes. The dose profiles are then aggregated across routes to construct an individual 1-year

profile. Exposure and dose metrics of interest (e.g., peak, time-averaged, time-integrated) are

extracted from the individual’s profiles, and the process is repeated thousands of times to obtain

population distributions. This approach allows identification of the relative importance of routes,

pathways, and model inputs. Two-stage Monte-Carlo sampling is applied to allow explicit

characterization of both variability and uncertainty in model inputs and outputs.

The outputs from the SHEDS-Pesticides model include: (1) the time profile of exposure

and dose metrics for a specified post-application time period for an individual; (2) exposure and

dose contribution by routes and pathways; (3) cumulative density function (CDF) and box-

7

Figure 2-1. Schematic Diagram of the SHEDS-Pesticides Simulation Process.

Characterize Variability and Uncertainty in Population

Read Scenario Information

Sample an Individual

Collect Pesticides Usage Information

Simulate Individual 1-year Route-Specific Daily Exposure Rate

Simulate Dose Over All Routes

Two-Stage Monte Carlo Simulation

Variability Run

Uncertainty Run

8

whisker graphs for aggregate population estimates and uncertainty of percentiles.

2.2 Main Characteristics of the SHEDS-Pesticide Model

The main characteristics of the SHEDS-Pesticides model include: (1) non-linearity and

interaction between inputs; (2) saturation points; (3) different input types (e.g., continuous versus

categorical); and (4) aggregation and carry-over effects.

Non-linearity is a relationship between two variables in which the change in one variable

is not simply proportional to the change in the other variable. An example in the SHEDS-

Pesticides model is an exponential decay term in the airborne concentration of the pollutant

(Equation A.3 in Appendix A).

Interaction is a case in which the effect of an input depends on the value of another input.

Interaction terms can be introduced in a model in multiplicative forms.

A saturation point is an input value above which the model output does not respond to

changes in the input. For example, there is an upper limit for possible dermal exposure via body

and hand in the SHEDS-Pesticides model (Equation A.8 in Appendix A).

Inputs may be qualitative (categorical) (e.g., gender) or quantitative (e.g., pesticide

concentration). Quantitative inputs can be continuous (e.g., body weight) or discrete (e.g.,

number of application of a pesticide). Some quantitative inputs may be described by empirical

distributions, while other inputs may be represented by parametric distributions.

Aggregation refers to situations in which multiple numerical values are combined into

one numerical value, such as sum or mean value. An example is the total exposure (ETotal) in the

SHEDS-Pesticides model that is an aggregate of exposure via inhalation (EInhalation), dermal

(EDermal), and ingestion (EIngestion) pathways.

9

Figure 2-2. Schematic Diagram for the Simplified SHEDS-Pesticides Model.

The carry-over effect is a situation in which the exposure for an individual in each day is

a function of exposure in the same day and the prior day(s). For example, dermal exposure via

hand is corrected for the carry-over effect (Equation A.11 in Appendix A).

2.3 Simplified Version of the SHEDS-Pesticides Model

A simplified version of the SHEDS-Pesticides model is developed for evaluation of

sensitivity analysis methods. Figure 2-2 shows a schematic diagram for the simplified SHEDS-

Pesticides model. The final exposure is an aggregate of possible exposures via three pathways

including inhalation, dermal, and ingestion. Key equations of the simplified model along with

probability distributions of model inputs are given in Appendix A.

Two key points are considered when developing the simplified model: (1) the simplified

model should have the key characteristics of the SHEDS-Pesticides model; and (2) the

simulation time for the simplified model should be realistically practical. The results from the

simplified model are not real exposures. However, these values can be used for evaluation of

10

selected sensitivity analysis methods. In order to address the key points, the following

assumptions are made:

• Because the main objective of this evaluation is to identify which model inputs are major

contributors to the variance of model outputs, the effect of specific human activity

patterns and application patterns of pesticides on model outputs are not considered.

• One-stage Monte Carlo simulation is considered for the probabilistic framework of the

simplified model with a focus on variability in inputs.

• Exposure duration is set for one month in order to reduce the simulation time.

• There is only one application of pesticides during exposure period. The application

occurs at the first day of exposure duration.

• Daily exposures from multiple exposure pathways are calculated based upon randomly

generated exposure times corresponding to pathways, rather than based upon the human

time-activity pattern database used by the SHEDS-Pesticides model. This will help

reduce the complexity of simulation process and the simulation time.

Table 2-1 provides a summary of example inputs to the simplified SHEDS-Pesticides model

including their probability distribution and sampling frequencies. A complete list of inputs is

provided in Appendix A.

Typically, inputs have two different sampling strategies within the time scale of the

model simulation (e.g., one month): (1) monthly; and (2) daily. Inputs with a monthly sampling

strategy are sampled only once for each individual within each month, while in contrast, inputs

with a daily sampling strategy are sampled every day for each individual within the time scale of

the model simulation.

11

Table 2-1. Examples for Input Assumptions for the Simplified SHEDS-Pesticides Model

Inputs Distribution (a) Unit Comment Sampling Frequency

DR Triangular (0.1, 0.2, 0.4) [-] Fraction of residue that dissipates daily

monthly

PWB Triangular (0.3, 1.0, 1.0) [-] Probability of washing body monthly

FTR Uniform (0.0, 0.08) [-] Fraction of chemical available for transfer daily

a For triangular distributions, minimum, most likely, and maximum values are given For uniform distributions, minimum and maximum values are given

Examples of inputs with a monthly sampling strategy include fraction of residue that

dissipate daily (DR) and probability of washing body (PWB). DR represents a decay rate for

chemical residue on a daily basis. PWB represents a probability that an individual washes his/her

body in each day, and hence, increasing the chance of removing the chemicals from the body

surface. The fraction of chemicals available for transfer (FTR) is an input with a daily sampling

strategy. These inputs are later used in Chapter 5 when examples are provided to illustrate the

results of each sensitivity analysis method. Typically, these inputs are selected among the list of

top sensitive inputs in the model by different sensitivity analysis methods.

13

3. OVERVIEW OF UNCERTAINTY AND SENSITIVITY ANALYSIS METHODS

The objective of this chapter is to briefly review typical methods for: (1) propagation of

probability distributions of inputs through a model (Section 3.1); and (2) sensitivity analysis of a

model (Section 3.2). More detailed discussion methods for propagating probability distributions

of inputs through a model and of sensitivity analysis methods is given in Volume 1. This chapter

provides a brief summary of these methods for the convenience of the reader. Sensitivity analysis

methods introduced in this chapter are further evaluated through a set of case studies in Chapter

5.

3.1 Methods for Propagation and Quantification of Variability and Uncertainty

In order to quantify variability and uncertainty, the distribution of model inputs should be

propagated through the model to obtain distributions on model outputs. Propagation techniques

may be analytical, approximation, or numerical.

3.1.1 Analytical Propagation Techniques

For simple models in which the output is a function of linear combination of model

inputs with no dependency, the propagation of uncertainty through the model is straightforward.

For example, considering the Central Limit Theorem, the mean and variance of the output are

equal to the sum of the means and variance of the inputs (Wilson and Crouch, 1981; DeGroot,

1986). The same approach can be used for models with outputs as functions of multiplicative

independent input using a logarithmic transformation. Although analytical propagation is

conceptually straightforward and easy to implement, its applicability is limited to models with

linear summation (or product) of independent inputs.

14

3.1.2 Approximation Methods

Typical approximation methods for propagation of uncertainty in a model are generation

of system moments and first order methods (Hahn and Shapiro, 1967). These methods typically

use Taylor series expansions to propagate the moments of input distributions (e.g., mean,

variance, skewness, and kurtosis) through a model to quantify the equivalent moments of the

output values. The distribution of the output values can be approximated with propagated

moments using parametric distributions. Approximation methods typically have three major

limitations (Cullen and Frey, 1999). These methods require differentiable functional form for the

model, and hence, they are not applicable to models with discrete or discontinuous behaviors.

Approximation methods are also computationally intensive. Moreover, information regarding the

tails of the input distributions cannot be propagated with approximation methods.

3.1.3 Numerical Propagation Techniques

For most real-world applications involving complex and nonlinear model, analytical

techniques and often approximation methods are not favorable. The most common techniques for

numerical propagation of uncertainty and variability are sampling based methods. Sampling

based methods do not require access to the model equations. These methods involve running a

set of model simulations at a set of sampled points from probability distributions of inputs and

establishing a relationship between inputs and outputs using the model results. Some of the

commonly used sampling based uncertainty analysis methods are: (1) Monte Carlo Methods; (2)

Latin Hypercube Sampling methods; (3) Fourier Amplitude Sensitivity Test (FAST); and (3)

reliability based methods.

Monte Carlo Simulation. Monte Carlo simulation (Rubinstein, 1981; Doll and

Freeman, 1986; Kalos and Whitlock, 1986; Sobol, 1994; Fishman, 1996) is widely used for

15

propagating probability distributions of inputs through a model. Monte Carlo simulation involves

random sampling from distribution of inputs and successive model simulation to estimate the

distribution for the model output. Monte Carlo simulation can be used to solve problems with

physical probabilistic structures, such as uncertainty propagation in models or solutions of

stochastic equations, or can be used to solve non-probabilistic problems, such as finding the area

under a curve (Gardiner, 1983; Devroye, 1986). The applicability of Monte Carlo simulation is

sometimes limited to simple models because this method typically requires large number of

samples. For computationally intensive models, the time and resources required for Monte Carlo

simulation can be prohibitive.

Latin Hypercube Sampling. LHS method (Iman and Conover, 1980; Stein, 1987; Loh,

1996) is a variant of the standard Monte Carlo simulation. In LHS the probability distribution for

the random variable of interest is first divided into ranges of equal probability, and one sample is

taken from each equal probability range. However, the order of the samples is typically random

over the course of the simulation, and the pairing of samples between two or more random inputs

is usually treated as independent (Cullen and Frey, 1999). LHS method typically provides a

better coverage of the entire range of the probability distribution of an input compared to Monte

Carlo simulation.

Fourier Amplitude Sensitivity Test (FAST). FAST has been developed for uncertainty

and sensitivity analysis (Cukier et al., 1973; 1975, 1978; Schaibly and Shuler, 1973). FAST

provides a way to estimate the expected value and variance of the output variable and the

contribution of individual factors to this variance (Saltelli et al., 2000). FAST is a method based

on Fourier transformation of uncertain model parameters into a frequency domain. Through

Fourier transformation, a multi-dimensional model will be reduced into a single dimensional one.

16

Reliability Based Methods (FORM and SORM). First and second-order reliability

methods, also known as FORM and SORM, estimate the probability of an event under

consideration. These methods can provide the probability that an output exceeds a specific value,

also known as probability of failure (Karamchandani and Cornel, 1992; Lu et al., 1994; Hamed

et al., 1995). These methods are useful in uncertainty analysis of models that have a single

failure criterion.

3.2 Sensitivity Analysis

A series of seven sensitivity analysis methods that are of practical significance were

suggested in Volume 1 for further evaluation. These methods include Pearson and Spearman

correlation analyses, sample and rank regression analyses, analysis of variance (ANOVA),

Sobol’s method, and Fourier amplitude sensitivity test (FAST). These methods are briefly

explained in the following sections. Chapter 5 presents the results from evaluation and

comparison of these methods based on application to the simplified SHEDS-Pesticides model.

3.2.1 Correlation Analysis

Correlation coefficients, which are the special cases of regression-based analysis, are

typically useful methods for sensitivity analysis (Saltelli et al., 2000). The Pearson correlation

coefficient, for instance, can be used to characterize the degree of linear relationship between the

output values and sampled values of individual inputs. If the relationship between an input and

an output is nonlinear but monotonic Spearman correlation coefficients based upon rank

transformed values of an input and an output provides better performance compared to Pearson

correlation coefficients (Gibbons 1985, Siegel and Castellan 1988, and Kendall 1990). Neither

Pearson nor Spearman correlation coefficients can provide insight regarding possible interaction

17

effects between inputs. Resulting correlation coefficients can be used to rank the input values

based upon the magnitude of their influence on the output.

3.2.2 Linear Regression Analysis

Linear regression-based approaches for sensitivity analysis have the advantage that they

can evaluate the influence of various model inputs simultaneously compared to correlation-based

techniques (Saltelli et al., 2000). When each input in the regression model is standardized such

that the inputs are normally distributed about zero with a standard deviation of one, the

regression coefficients can provide insight into the relative significance of inputs (Devore and

Peck, 1996; Neter et al., 1996). Linear regression can also be performed using rank transformed

values of data instead of true values resulting in determination of rank regression coefficients. A

major limitation of regression-based approaches is their assumption of a known relationship

between model inputs and the output (e.g., linear, monotonically nonlinear, pre-specified

nonlinear terms, etc).

3.2.3 Analysis of Variance

If the relationships between the output and inputs cannot be precisely encoded into the

regression equations, there is a need for methods that do not assume any functional form for such

a relationship. One such approach is multi-factor ANOVA (Archer et al., 1997; Frey and Patil,

2002). ANOVA can quantify the relationship between the variation in one or more model inputs

and the changes in the mean of a model output (Neter et al., 1996). To perform ANOVA, the

model output values are binned as a function of their corresponding inputs. The influence of the

inputs can be evaluated by comparing the mean output value for each bin with that of the entire

dataset (Krishnaiah, 1981). ANOVA uses F tests to evaluate the statistical significance of input-

output relationship.

18

3.2.4 Sobol’s Method

Sobol’s method (Sobol, 1993; Saltelli et al., 2000) is a variance-based sensitivity analysis

technique based on Total Sensitivity Indices (TSI) that takes into account interaction effects

between inputs. The TSI of an input is defined as the sum of all of sensitivity indices involving

that input. The TSI involves first order effects as well as those of higher orders. Sobol’s method

is implemented by decomposing the input-output relationship into a summand of increasing

dimensionality. Solving the resulting equation to obtain sensitivity indices requires the use of

numerical integration techniques such as Monte Carlo, and is thus typically highly

computationally intensive. The total effects of inputs based on TSI do not provide a complete

characterization of the sensitivity. The interaction effect of an input (xi) with a specific input

(e.g., xi × xj) cannot be inferred from the TSI of the input (xi). However, the total effects are

much more reliable than the first-order effects in order to investigate the overall effect of each

single input on the output. Sobol’s method is model independent and can identify the

contribution of individual inputs to the output variance. However, Sobol’s method, in general, is

computationally intensive.

Sobol’s method is not readily available in commonly used statistical software packages.

Thus, an algorithm was prepared and coded in Matlab for application of Sobol’s method to the

case studies with the simplified SHEDS-Pesticides model. The algorithm is briefly explained in

Appendix B. This work represents the first time that Sobol’s method has been applied to a model

with multiple simulation time periods for multiple inputs. Thus, the basic algorithm prepared is

improved for application to the simplified SHEDS-Pesticides model with alternative sampling

strategies for inputs. The basic algorithm was verified based on results available from literature

(e.g., Saltelli 2002).

19

The test case used for verification of the Sobol’s algorithm was an analytic function

known as “Sobol’s g function”. This example was chosen because the cost of computation was

negligible and analytical answers were available for comparison with results based on Sobol’s

method.

The g function is a strongly non-monotonic, non-aditive function of k factors, xi, assumed

identically and uniformly distributed. The functional form of the “g function” can be given as:

∏= +

+−×=

k

i i

iik a

axxxxg

121 1

|24|),...,,( (3-1)

The importance of each input, xi, is driven by its associate coefficient ai. For ai = 0, the input is

important ( 2)(0 ≤≤ ii xg ). For, e.g., ai= 9 the input is non-important ( 1.1)(9.0 ≤≤ ii xg ), while

ai= 99 for the input can be considered as non-influent ( 01.1)(99.0 ≤≤ ii xg ). Further explanation

of the verification case study is given in Saltelli (2002).

The algorithm prepared for Sobol’s method was able to appropriately reproduce the

results given in Saltelli (2002). However, in some preliminary analyses with a simple linear

model that included the scale effect (i.e., inputs with substantially different mean values), the

algorithm showed slow convergence to the true sensitivity coefficients. However, when the

outputs from the algorithm were normalized, it was found that accurate results could be obtained

with a much smaller number of samples than implied by the raw results. The output values were

normalized using mean and standard deviations as:

y

yyyσ

)( −=′ (3-2)

where, y is the output value, y is the mean of the output values, and σy represents the standard

deviation of the output values.

20

In practice with our case studies with the simplified SHEDS-Pesticides model, the

algorithm developed for the Sobol’s method provided reasonable results in comparison to other

sensitivity analysis methods (See Chapter 5). Thus, a convergence problem is not suspected here.

3.2.5 Fourier Amplitude Sensitivity Test (FAST)

FAST (Cukier et al, 1973; Saltelli et al., 2000) is another variance-based global

sensitivity and uncertainty analysis procedure. FAST provides a way to estimate the expected

value and variance of the output variable and the contribution of individual inputs to this

variance. Similar to Sobol’s method, FAST provides first order indices representing the main

effects of inputs and TSI for model inputs. FAST is independent of the model structure. The

main difference between FAST and Sobol’s method is the approach by which the multi-

dimensional integrals are calculated. Whereas Sobol’s method uses a Monte Carlo integration

procedure, FAST uses a pattern search based on a sinusoidal function. The key advantage of

FAST is its capability in apportionment of the output variance to individual inputs in the model.

However, FAST is computationally complex for a model with a large number of inputs.

Similar to Sobol’s method, FAST is not readily available in commonly used statistical

software packages with exception of Simlab (Saltelli et al., 2004). This work represents the first

time that FAST has been applied to a model with multiple simulation time periods for multiple

inputs, and thus the case studies reported here is at the frontier of applicability of FAST.

Therefore, a modified algorithm for application of FAST was developed and coded in Matlab.

The most significant difference in this application of FAST versus those reported in the literature

is the need to deal with a much larger number of inputs.

Because the simplified SHEDS-Pesticides model has 30 time periods for each of the daily

inputs, and because the goal was to generate independent samples in each of the time periods,

21

there is a need for different frequencies for a given input variable for each time period (day) of

the simulation. Thus, although there are only 14 daily inputs, there is a need to generate 30 × 14

(i.e., 420) independent sets of samples. In addition, there are 11 monthly inputs. Thus, the total

number of independently sampled distributions is 431. Thus, there was a need to generate

additional frequencies beyond the set of 50 incommensurate frequencies available in literature.

This algorithm is briefly discussed in Appendix B. Basic algorithm of FAST was verified with

the same set of test cases used for verification of Sobol’s method. The algorithm prepared for

FAST was able to appropriately reproduce the results given in Saltelli (2002).

This chapter has briefly summarized methods that are discussed in more detail in Volume

1. Because the focus of this work is on evaluation of methods for sensitivity analysis, and

because there are not readily available software packages for Sobols’ and FAST methods,

algorithms were developed and coded for these two methods based upon information in the

literature. In the case of FAST, the algorithm had to be modified in a manner that allowed for

more inputs than the limit of 50 that has been used for years, but possibly could introduce a

dependency structure in the simulation that might compromise the case study results. However,

in order to make a preliminary assessment of whether FAST could be appropriate in the future,

the current version of the modified algorithm was applied to the simplified SHEDS model test

bed, along with the other methods.

23

4. DEFINING CASE SCENARIOS FOR SENSITIVITY ANALYSIS

The objective of this chapter is to define a case study scenario as the basis for performing

sensitivity analysis on the simplified SHEDS-Pesticides model. A scenario is a set of

assumptions about the nature of the problem to be analyzed (Cullen and Frey, 1999). These

assumptions may be based upon recommendations from field experts, stakeholders, risk

managers, or combinations of all three.

The definition of the most relevant or important scenarios is especially crucial in

situations for which there are limitations of time and other resources with respect to performing

sensitivity analysis. Thus, it is important to identify scenarios that are the highest priority for

evaluation. A well defined case scenario will help concentrate the sensitivity analysis on the

areas that are of more interest to risk managers and decision makers.

Figure 4-1 shows a schematic diagram of the case study scenario components for

sensitivity analysis of the simplified SHEDS-Pesticides model. This figure shows that issues

such as susceptible subpopulations, pathways of exposure, and time scale of the model

simulation should be clearly specified in the case scenario. Furthermore, the probabilistic

dimensions that are intended to be the focus of sensitivity analysis should be selected in a case

scenario. The methodology for sensitivity analysis is not dependent on these factors. Thus, for

purpose of demonstrating methods, it is not necessary to consider all possible scenarios.

Furthermore, in order to have meaningful outcomes from the sensitivity analysis that can be used

by risk managers, it is useful to define specific case scenarios relevant to the model scope that

are of policy interest.

24

Figure 4-1. Components of a Scenario for Sensitivity Analysis of the Simplified SHEDS-

Pesticides Model.

In the following sections, the major components of the scenario defined for sensitivity

analysis are explained. These components are classified into four categories including: (1)

probabilistic dimensions; (2) susceptible subpopulation; (3) identification of pathways of

interest; and (4) time scales of the model simulation. Sections 4.1 through 4.4 provide

discussions for these topics, respectively. For the defined scenario, three ways of summarizing

the results as a function of time are considered. Thus, there are multiple scenarios with respect to

interpretation of the temporal aspects. However, these scenarios are based upon one set of

assumptions regarding the other aspects.

4.1 Probabilistic Dimensions

The probabilistic dimension of a scenario indicates whether the assessment will explicitly

incorporate variability, uncertainty, or both. A scenario will typically include variability in

exposures among different members of a population, unless the scenario is for a single

individual. Regardless of whether the scenario is for a population or for a single individual, there

will typically be uncertainty in the inputs to a model. Therefore, as a part of defining the

scenario, it is important to define the probabilistic dimension of the scenario. Furthermore, there

Scenario

Probabilistic Dimension

Susceptible Subpopulation

Time Scale

Exposure Pathway

25

are alternative methods for dealing with variability and uncertainty. The choice of an appropriate

method should be made taking into account the assessment objectives, the data quality

objectives, the availability of data, and the importance of the assessment. Because the simplified

SHEDS-Pesticides model has a one-dimensional probabilistic framework considering variability

in inputs, variability-only analysis is selected for the probabilistic dimension of sensitivity

analysis.

The purpose of an analysis of variability-only is to quantify inter-individual variability in

exposure and risk. Such an analysis is typically predicated on the assumption that the range of

variability is much larger than the range of uncertainty; therefore, a judgment is made that

uncertainty can be neglected. The appropriateness of this assumption will depend upon the

specific problem and the objectives of the analysis. Variability can include controllable or

explainable sources of variation (e.g., differences in time spent in pesticide contaminated area

among members of a population) or stochastic sources of variability (e.g., differences in fraction

of chemical available for transfer on a daily basis).

4.2 Susceptible Subpopulation

The scope of the risk assessment is typically on specific population groups in which the

risk of adverse effect due to exposure to the hazard is expected to be significant. For a particular

pollutant, there may be multiple susceptible subpopulations because different groups of people

may be exposed to that particular type of pollutant. Susceptible subpopulations can be

incorporated in the model using alternative dose-response relationships. Furthermore, exposed

subpopulations could be differentiated by age classes with respect to activity patterns, weight,

and inhalation rate.

26

For the purpose of evaluation of selected sensitivity analysis method, children between 5

and 10 are considered as the population of interest in the simplified SHEDS-Pesticides model.

Typically, children between 5 and 10 spend considerable amount of time outdoor in play

grounds, and hence, have high potentials for exposure to pesticides. The information regarding

body height and weight for each member of population is provided with consideration of the

selected age range. As a caveat, because sensitivity analysis is therefore focused on a subset of

the general population, insights regarding key sources of variability may not be representative of

other groups.

4.3 Identification of Pathway of Interest

In order to adequately characterize total exposures, a scenario may need to consider

multiple pathways of exposure. The simplified SHEDS-Pesticides model considers three

different exposure pathways including inhalation, dermal (via body and hand), and ingestion.

Total exposure represents the aggregation of exposure from all three pathways. Figure 4-2 shows

probability distribution of average exposure from each of the three pathways and for the average

total exposure in 30 days of model simulation. The results of the different exposure pathways in

this figure are based upon the simplified SHEDS-Pesticide model and input assumptions

explained in Chapter 2 and Appendix A. The ranking of three exposure pathways with respect to

the magnitude of the average monthly exposure is: (1) dermal; (2) ingestion; and (3) inhalation

pathway. For the purpose of sensitivity analysis and evaluation of selected methods, total

exposure is considered as the output of interest.

4.4 Time Scales of the Model Simulation

A risk assessment scenario may include a temporal dimension. The temporal

considerations for a scenario typically include: (1) the time for each major step in the exposure

27

0

0.2

0.4

0.6

0.8

1

0.001 0.1 10 1000

Mean Exposure in 30 Days (µg/kg)

Cum

ulat

ive

Pro

babi

lityInhalation ExposureIngestion ExposureDermal ExposureTotal Exposure

Figure 4-2. Probability Distribution of Different Average Exposure Pathways in 30 Days of Model Simulation.

episode, in order to estimate the exposure at each step; (2) the activity patterns of individuals

with regard to frequency of exposure to particular types of pollutant; (3) “temporal dynamics”

effects, whether at a short time scale (e.g., daily, weekly) or a longer scale (e.g., monthly,

seasonal, annual); and (4) the time period associated with occurrence of illness as a result of one

or more exposures (Frey et al., 2004). In the simplified SHEDS-Pesticides model, inputs have

different sampling frequencies with respect to time. Some inputs such as body weight or amount

of applied pesticide are only sampled once for each individual during the entire simulation time

of the model (e.g., 30 days). In contrast, some other inputs such as inhalation time and body

washing removal efficiency are sampled in shorter time steps for each individual (i.e., daily).

Three analysis scenarios are defined in order to evaluate the implications of different time scales,

including: (1) Scenario I - the daily total exposures; (2) Scenario II - incremental change in daily

exposure representing the rate of change of exposures from one day to the next; and (3) and

Scenario III - longer-term cumulative exposures.

28

Figure 4-3 shows variation of total exposure for selected random individuals in 30 days

of model simulation. The results are based upon the simplified SHEDS-Pesticide model and

input assumptions explained in Chapter 2 and Appendix A. Daily exposure for each individual is

the direct output from the model at each time step of the simulation, )(ty , where y is the total

exposure from the model and t represents the time step of the model simulation. Incremental

change in daily exposure represents the difference between the exposure at the current time step

and the previous time step (i.e., )1()( −− tt yy ). Cumulative exposure is sum of exposures from day

1 to the current time step of the model simulation (i.e.,∑=

t

iiy

1)( ).

Table 4-1 summarizes an example of the output of the model for the three time scales of

the scenario. Results are given for the individual ‘a’ shown in Figure 4-3. Incremental change in

daily total exposure was either negative or positive at different time steps of the model

simulation, while daily and cumulative exposures were always positive. For example, because

daily total exposures in the first and second days were 48.9 and 25.1 µg/m3, respectively, the

incremental change in daily total exposure in the second day had a negative value of -23.8

µg/m3. In contrast, in the third day, the incremental change in daily total exposure had a positive

value of 13.8 µg/m3 due to increase in the daily exposure from 25.1 to 38.9 µg/m3 between the

second and third days, respectively.

29

(a) Scenario I: Daily Exposure

0

20

40

60

80

100

1 6 11 16 21 26 31

Time (day)

Tota

l Exp

osur

e ( µ

g/kg

)Individual a Individual b Individual c Individual d

(b) Scenario II: Incremental Change in Daily Exposure

-60

-40

-20

0

20

40

60

1 6 11 16 21 26 31

Time (day)

Tota

l Exp

osur

e ( µ

g/kg

)

(c) Scenario III: Cumulative Exposure

0

100

200

300

400

500

600

700

800

1 6 11 16 21 26 31

Time (day)

Exp

osur

e ( µ

g/kg

)

Figure 4-3. Example of Probabilistic Results for Variation in the Total Exposure for Selected Random Individuals: (a) Scenario I-Daily Exposure; (b) Scenario II-Incremental Change in

Daily Exposure; and (c) Scenario III-Cumulative Exposure.

30

Table 4-1. Example of Probabilistic Results for Daily, Incremental Change in Daily, and Cumulative Total Exposure for the Random Individual ‘a’ shown in Figure 4-3

Time [day]

Scenario I: Daily Total Exposure

[µg/m3]

Scenario II: Incremental Change in Daily

Total Exposure [µg/m3]

Scenario III: Cumulative Total

Exposure [µg/m3]

1 48.9 48.9 (a) 48.9 2 25.1 -23.8 74.0 3 38.9 13.8 112.9 4 40.9 2.0 153.8 5 77.2 36.3 231.1 6 61.9 -15.3 293.0 7 32.8 -29.1 325.8 8 48.5 15.7 374.3 9 35.1 -13.4 409.4 10 14.2 -20.9 423.6 11 23.4 9.2 447 12 20.4 -3.0 467.4 13 28.0 7.6 495.4 14 27.0 -1.0 522.4 15 20.5 -6.5 542.9 16 23.8 3.3 566.7 17 29.0 5.2 595.7 18 21.6 -7.4 617.3 19 21.5 -0.1 638.8 20 14.8 -6.7 653.6 21 10.0 -4.8 663.6 22 11.1 1.1 674.7 23 8.6 -2.5 683.3 24 6.8 -1.8 690.1 25 3.0 -3.8 693.1 26 3.3 0.3 696.4 27 4.8 1.5 701.2 28 6.4 1.6 707.6 29 3.0 -3.4 710.6 30 2.0 -1.0 712.6

a The incremental change in daily total exposure at Day 1 is assumed to be equal to the total daily exposure in that day

31

5. RESULTS OF SENSITIVITY ANALYSIS FOR CASE STUDIES

The objective of this chapter is to present the results from applying selected sensitivity

analysis methods to the case scenario defined in Chapter 4. Section 5.1 provides a brief

discussion regarding the model application, including general algorithm used for application of

the selected sensitivity analysis methods, three different temporal scenarios considered in the

analysis, and the sample size for methods based on Monte Carlo simulation. The specific

methods evaluated here include Pearson and Spearman correlation analyses (Section 5.2), linear

sample-based and rank-based regression (Section 5.3), ANOVA (Section 5.4), Sobol’s method

(Section 5.5), and FAST (Section 5.6). These methods are briefly discussed in Chapter 3 and

extensively explained in Volume 1. Section 5.7 presents a comparison of results based on

different sensitivity analysis methods.

5.1 Model Application

The simplified SHEDS model documented in Chapter 3 was used as a testbed for

development of quantitative sensitivity analysis results for each of the seven selected sensitivity

analysis methods. As described in Chapter 4, the simplified model is applied to an exposure

scenario involving residential exposure to pesticides. The pesticide is applied at the beginning of

a month and exposures are estimated on both a daily and cumulative basis for all exposure

pathways.

The application of the seven selected sensitivity analysis methods involves sensitivity

analysis for each time step (day) of the model simulation. For the six sampling-based methods,

which include all of the selected methods except FAST, probability distributions representing

inter-individual variability in the inputs were propagated through the simplified SHEDS-

Pesticide model using Monte Carlo simulations with a sample size of 10,000. For inputs that had

32

a monthly sampling strategy, and hence, did not vary on a day-to-day basis, one value per

individual per month was sampled, for a total of 10,000 samples. These values were used for all

time steps of the model simulation. In contrast, for inputs that had a daily sampling strategy, and

hence, varied on a day-today basis, 30 values per individual were sampled with a different value

in each time step of the model simulation. Thus, there are a total of 300,000 samples for those

inputs that vary daily and also among individuals.

The sample size for FAST was assigned based on the number of inputs and the maximum

frequency associated with inputs. Further explanations for selecting the sample size for FAST are

given in Volume 1.

Figure 5-1 shows a schematic diagram of the general algorithm for application of the

selected sensitivity analysis methods to the simplified SHEDS-Pesticide model. Each dataset in

this figure includes randomly sampled values from variability distributions of inputs at a given

day for sampling-based methods or values generated from domains of inputs using the

transformation functions in FAST and the corresponding model output values. Output values

were based on three temporal scenarios introduced in Chapter 4, which are: (1) Scenario I - daily

total exposure; (2) Scenario II - incremental change in daily total exposure; and (3) Scenario III -

cumulative total exposure. Sensitivity analysis was applied separately to each dataset. At each

time step, the key sources of variability were identified. This process was repeated n times to

arrive at different sensitivity rankings for inputs in which n refers to the total number of time

steps considered in the model simulation (i.e., 30 days). A ranking represents the comparative

order of importance of each input on a given day when the inputs are sorted according to their

sensitivity indices. A rank of 1 refers to the most important input. As the numerical value of the

33

Figure 5-1. Schematic Algorithm for Application of Selected Sensitivity Analysis Methods to the Simplified SHEDS-Pesticides Model.

rank increases, the importance decreases, the rank is said to be worse. The variation of rankings

when comparing different time steps were used to assess whether the sensitivity of inputs

changed at different times based on different types of model responses to each input and on

possible differences in the importance of inputs with sampling strategies of monthly versus daily

at different times of the month.

5.2 Correlation Analysis

This section presents the results based upon the use of Pearson and Spearman correlation

coefficients to characterize the sensitivity of the selected model outputs for temporal Scenarios I,

II, and III to each of the inputs of the model. For discussion purposes, this section focuses on

providing specific details for selected inputs, followed by an overall summary of results for all of

the model inputs. The results presented in this section include: Pearson correlation coefficients

(PCC) and Spearman correlation coefficients (SCC) for selected inputs; changes in rankings of

Dataset 1

Dataset 2

Dataset n

Sensitivity Analysis

Sensitivity Analysis

Sensitivity Analysis

Ranking 1

Ranking 2

Ranking n

Summary of Sensitivity

Analysis Results

n : Number of time steps for model simulation (e.g., n=30)

34

selected inputs with respect to time; and mean ranks and ranges of ranks during the 30 days of

model simulation for all inputs. The inputs selected for detailed discussion include residue decay

rate (DR), probability of washing body (PWB), and fraction of chemicals available for transfer

(FTR). These inputs are further explained in Section 2.3. These inputs were selected because they

illustrate the following key issues: (1) they have different sampling strategies; (2) they have

different patterns with respect to sensitivity indices and rankings with respect to time; and (3)

typically, they were among the most sensitive inputs.

5.2.1 Correlation Coefficients Results

Regardless of the frequency with which inputs were sampled, correlation coefficients

were estimated at each day for all inputs for each of the three temporal scenarios, and hence, 30

sets of Pearson and Spearman correlation coefficients were estimated. This is because there

could be changes over time in sensitivity of inputs sampled on a monthly basis depending on the

sensitivity of phenomena that occurred on a daily basis. Typically, both Pearson and Spearman

correlation coefficients for each input varied in time in the three temporal scenarios. For inputs

with a monthly sampling strategy, two distinct patterns with respect to the change in the absolute

value of the correlation coefficients over time were identified: (1) monotonic decrease; and (2)

monotonic increase. However, for inputs with a daily sampling strategy, correlation coefficients

typically declined in time. Examples are provided to illustrate and explain these patterns.

Figure 5-2 shows variation in absolute values of PCC and SCC and corresponding 95%

confidence intervals for the residue decay rate (DR) with respect to time. The coefficients are

shown for the three temporal scenarios. DR has a monthly sampling strategy and provides an

example in which the input tends to have a higher correlation coefficient later in the month.

35

(b) Scenario II: Incremental Change in Daily Exposure

0

0.25

0.5

0.75

1

0 5 10 15 20 25 30

Time (day)

|Cor

rela

tion

Coe

ffici

ent|

PearsonSpearman

(c) Scenario III: Cumulative Exposure

0

0.25

0.5

0.75

1

0 5 10 15 20 25 30

Time (day)

|Cor

rela

tion

Coe

ffici

ent|

PearsonSpearman

Figure 5-2. Absolute Values and Corresponding 95% Confidence Intervals of Pearson and Spearman Correlation Coefficients for Re sidue Decay Rate as an Input with Monthly Sampling

Frequency for: (a) Scenario I: Daily Exposure; (b) Scenario II: Incremental Change in Daily Exposure; and (c) Scenario III: Cumulative Daily Exposure.

(a) Scenario I: Daily Exposure

0

0.25

0.5

0.75

1

0 5 10 15 20 25 30

Time (day)

|Cor

rela

tion

Coe

ffici

ent|

PearsonSpearman

95% Confidence Interval of the Coefficient

36

Figure 5-2 (a) shows the results for Scenario I. DR has a substantially higher correlation with

daily total exposure later in the month. For example, SCC and PCC had low values of 0.08 and

0.05, respectively, in the first day. However, these coefficients increased to 0.87 and 0.37,

respectively, by the end of the month. Increase in the magnitude of correlation coefficients, and

hence, the sensitivity of DR in later time steps of the model simulation can be attributed to the

functional form of the simplified SHEDS-Pesticides model. DR is used in the model when

chemical residues at target and non-target surfaces are estimated on a daily basis (See Equations

13-a and 13-b in Appendix A). However, because the effect of DR is raised to the power i, where

i represents the day for which the effect of residue decay rate is estimated, DR will have higher

contribution to variation of the model output later in the month and with the increase in the value

of i. Thus, although daily exposures were higher within the first few days of the month (See

Figure 4-3), DR was more correlated to the model output later in the month. However, because

the daily exposures typically decrease as more time elapses from the time of pesticide

application, the higher daily correlation late in the month for DR may not be of as much

significance when compared with other inputs that have high correlations earlier in the month,

when total daily exposures are higher.

Figure 5-2 (b) show the results for Scenario II for the change in PCC and SCC of DR with

respect to the time. Typically, the change in magnitude of correlation coefficients for this

temporal scenario follows a similar pattern compared to Scenario I. However, DR typically had

lower correlation with the incremental change in the daily exposures than with the total daily

exposures. The likely reason is that although the residue decay rate implies that there is a

continuous change over time in the amount of pesticide residue, the numerical value of DR itself

does not change from one day to the next. Thus, any value of DR other than zero will imply

37

some incremental change from one day to the next, but other inputs (and particularly those that

vary on a daily basis) are responsible for larger incremental changes. As noted in Chapter 4, and

in the example of Table 4-1, incremental changes can be positive or negative, and thus are

inherently “noisier” than the results from the other two temporal scenarios. For example, the

cumulative exposure in Scenario III must always increase monotonically, and typically the

marginal increase becomes smaller with time, and thus Scenario III would typically be expected

to produce the most stable results for sensitivity indices when comparing one day to the next,

whereas Scenario II would be expected to produce the least stable results.

Figure 5-2 (c) shows the results for Scenario III. Absolute values of PCC and SCC for DR

along with their 95% confidence intervals are shown in 30 days of model simulation. Similar to

the previous temporal scenarios, the magnitude of correlation coefficients increased in later time

steps of the model simulations. However, the magnitude of coefficients implies that DR had

lower correlation with the cumulative exposure at each day compared to the daily exposure in the

same day. The cumulative exposure for each individual at a selected day not only depends on the

value of DR in the same day, but also it relies on the values that DR held on previous days. As

shown in Figure 5-2(a), the daily exposures have a small correlation with DR early in the month.

However, especially in the first several days, the magnitude of the daily exposures is typically

the largest. By mid month, most of the cumulative exposure has occurred, and the marginal

increase in cumulative exposure becomes smaller, on average, on a daily basis. Thus, although

DR is highly correlated with the decreasing values of daily exposure in the later days of the

month, these daily exposures contribute relatively little to the cumulative total. For example, by

the 20th day, there is little perceptible increase in the magnitude of the correlation coefficients

38

shown in Figure 5-2(c), which is indicative of little or no marginal impact of DR late in the

month.

Comparing Pearson and Spearman coefficients for the three temporal scenarios indicates

that rank transforming the data typically increased the magnitude of the correlation coefficients

for DR. The highest increase in the magnitude of the coefficients was for Scenario I. DR has a

nonlinear relationship with daily exposure values. Because of the nonlinearity, the PCC is not a

good measure of sensitivity. The relatively low values of 0.30 for the PCCs in the last half of the

month do not appropriately quantify the strength of the relationship between daily exposure and

the selected input. However, although nonlinear, the relationship between daily exposure and DR

is monotonic. Therefore, the SCC provides a much better measure of the sensitivity of the output

to DR. as reflected in the larger magnitude of the correlation that is achieved based upon rank

transformation of the data.

For Scenario II, rank transforming the data did not improve the correlation between the

output values and DR. When the correlation is small for PCC, it is also small for SCC. The

differences between the two only become apparent when the correlations become larger.

Because Scenario II generally has low correlations, it matters less as to what type of correlation

coefficient is used to quantify the magnitude of the correlation. In other words, both methods

seem equally capable of indicating situations in which the correlations are very low.

For Scenario III, rank transforming data had a minimal effect on improving the

magnitude of correlation coefficients. The cumulative exposure is based upon the summation of

exposures over a given number days, and thus has a strong linear (additive) aspect. Although the

daily estimate of exposure has a nonlinear relationship with DR, the cumulative exposure is an

additive combination of multiple days. Furthermore, the cumulative exposure is most sensitive

39

to those days in which the highest magnitude of daily exposures occurred. The highest daily

exposures were early in the month. Early in the month, the residue decay rate was typically not

high enough to cause a large change in the amount of pesticide residue, and therefore the

behavior of the model of daily exposures to DR is approximately linear, especially in the first five

days. This is corroborated by the similarity of the PCC and SCC values for daily exposures in

the first five days. By the 10th to 15th day, the cumulative exposure is quite large compared to

that on the 30th day, and thus is most highly influenced by the effect of residue decay rate in the

first half of the month. Between the 5th and 15th day, the daily exposure response to DR becomes

increasingly nonlinear, which causes some nonlinearity in the response of the cumulative

exposure as well. However, the effect of the nonlinearity in cumulative exposure in the middle

part of the month is somewhat attenuated because of the stronger influence of the more linear

effect in the first five days.

Confidence intervals given for the correlation coefficients provide insight with respect to

statistically significant differences between coefficients. Typically, correlation coefficients based

on Pearson and Spearman methods did not statistically differ in the first few days of the model

simulation for the three temporal scenarios. However, typically later in the month, there were

statistically significant differences between PCC and SCC, typically because of nonlinearities in

the model as previously discussed. Furthermore, the confidence intervals can be used to assess

whether changes from one time period to another are significant. For example, for Scenario I,

there is a significance difference in sensitivity on the 30th day versus the 15th day based on SCCs,

but the SCCs during the last five days are approximately the same. Likewise, the SCCs are

approximately the same for the last 15 days for Scenario III.

40

Figure 5-3 shows variation in absolute values of PCC and SCC and corresponding 95%

confidence intervals for probability of washing body (PWB) with respect to time. PWB has a

monthly sampling strategy and provides an example in which the highest correlations for

Scenario I occurs early in the month. The magnitude of the correlation coefficients is relatively

modest, ranging from approximately 0.1 to 0.3. However, the association is significantly

stronger at the 5th day when compared to later in the month, such as after the 15th day. The

reason for a peak in sensitivity is that as time elapses, there is a change in the order of

importance among several inputs. For example, in the first few days of the month, the residue

decay rate is not one of the most important inputs. However, as time elapses, the variation in the

amount of pesticide residue remaining has a stronger relationship to residue decay rate. Thus, as

one input increases in importance, and therefore changes in rank toward ranks closer to one,

other inputs must simultaneously be displaced in terms of their importance. Thus, there is a

temporal dynamic in sensitivity.

The larger magnitude of the SCC compared to the PCC suggests that there is a moderate

nonlinear but monotonic association with this input. Furthermore, when compared to DR, PWB

has a higher magnitude of the SCC in the first several days, and thus has a more sensitive

ranking early in the month, when daily exposures tend to be the highest. For example, PWB has a

rank of one, whereas DR has a rank of six on Day 5. Over time, these rankings change, with DR

having an improved rank (closer to 1) later in the month, and PWB simultaneously having a

worsening rank. Of course, as noted earlier, the magnitude of daily exposures is highest early in

the month, and thus it is more significant that PWB is highly ranked (closest to 1) early in the

month.

41

(b) Scenario II: Incremental Change in Daily Exposure

0

0.2

0.4

0.6

0 5 10 15 20 25 30

Time (day)

|Cor

rela

tion

Coe

ffici

ent|

PearsonSpearman

(c) Scenario III: Cumulative Exposure

0

0.2

0.4

0.6

0 5 10 15 20 25 30

Time (day)

|Cor

rela

tion

Coe

ffici

ent|

PearsonSpearman

Figure 5-3. Absolute Values and Corresponding 95% Confidence Intervals of Pearson and

Spearman Correlation Coefficients for Probability of Washing Body as an Input with Monthly Sampling Frequency for: (a) Scenario I: Daily Exposure; (b) Scenario II: Incremental Change in

Daily Exposure; and (c) Scenario III: Cumulative Daily Exposure.

(a) Scenario I: Daily Exposure

0

0.2

0.4

0.6

0 5 10 15 20 25 30

Time (day)

|Cor

rela

tion

Coe

ffici

ent|

PearsonSpearman

95% Confidence Interval of the Coefficient

42

Early in the month, when PWB is one of the most important inputs with respect to daily

exposure, it tends to have a higher sensitivity with respect to the incremental change in daily

exposures, as shown in Figure 5-3(b). However, within the first few days, the magnitude of both

the SCC and PCC decreases significantly, and toward the end of the month these coefficients are

either not statistically significant or they are not of any practical significance. In contrast, as will

be shown in the next example, inputs that vary on a daily basis tend to have more influence on

the incremental amount of change from one day to the next late in the month, but not necessarily

on the late-in-the-month magnitude of the daily exposures or especially on the cumulative

exposures. Thus, the results for inputs that have a monthly sampling frequency with respect to

Scenario II are based in part on a competition with those inputs that vary on a daily frequency.

PWB increases in importance with respect to cumulative exposure in Scenario III, as

shown in Figure 5-3(c), during the first five days of the month. However, from Day 5 to Day 30,

there is no significant change in either the PCC or the SCC. This result is attributable to two

concurrent causes: (1) this input becomes less important on a daily basis later in the month; and

(2) the daily exposures become much smaller later in the month. Compared to DR, the main

difference is that PWB decreases in importance on a daily basis, whereas DR increased in

importance on a daily basis, most notably in days 5 to 10. Therefore, DR had some impact on an

increased cumulative sensitivity during that time. In contrast, PWB simultaneously declined in

importance as the magnitude of daily exposures decreased, and its impact on cumulative

exposure essentially remains “frozen” after the fifth day.

Typically, correlation coefficients based on the ranked transformed data were larger than

those based on the sampled data for Scenarios I and III, and there were small but statistically

significant differences especially for Scenario III. The similarity in the two types of correlation

43

coefficients may imply that if there is a nonlinearity effect, or some other type of interaction

effect, it is relatively small. For Scenario II, there was typically no significant difference

between the two types of correlation coefficients, but the magnitudes were typically small for

both.

Figure 5-4 shows an example of change in correlation coefficients for the fraction of

chemicals available for transfer (FTR). This input has a daily sampling frequency. For Scenario

I, the correlations were significantly higher on the first day than on the second day, and the

magnitudes of both the SCC and PCC decreased approximately monotically with time. The

reason for the declining importance of this input, and of other daily inputs not shown here, is

further described later in this section.

Although FTR rapidly declined as an influence on the magnitude of the daily exposure as

time elapsed, it remains one of the more important reasons for incremental changes in exposure

from one day to the next. Unlike the inputs that are sampled only once per month, the value of

FTR changes each day, which in turn can cause incremental changes in the daily exposures. As

illustrated in Table 4-1, the incremental changes in daily exposures tend to decrease as time

elapses. Thus, it is possible that an input could be highly important with respect to incremental

changes, but be of less importance to inter-individual variation in the magnitude of the daily

exposure. This is because some of the variation in the daily exposure magnitude on a given day

may be approximately unchanged from one day to the next, such as because of the influence of

inputs with a monthly sampling frequency. Typically, it is expected that inputs that vary with a

high frequency would be more important with respect to short-term incremental changes but

would not necessarily have a strong impact on the underlying causes of variability from one day

to the next.

44

(b) Scenario II: Incremental Change In Daily Exposure

0

0.2

0.4

0.6

0 5 10 15 20 25 30

Time (day)

|Cor

rela

tion

Coe

ffici

ent|

PearsonSpearman

(c) Scenario III: Cumulative Exposure

0

0.2

0.4

0.6

0 5 10 15 20 25 30

Time (day)

|Cor

rela

tion

Coe

ffici

ent|

PearsonSpearman

Figure 5-4. Absolute Values and Corresponding 95% Confidence Intervals of Pearson and

Spearman Correlation Coefficients for Fraction of Chemicals Available for Transfer as an Input with a Daily Sampling Strategy for: (a) Scenario I: Daily Exposure; (b) Scenario II: Incremental

Change in Daily Exposure; and (c) Scenario III: Cumulative Daily Exposure.

(a) Scenario I: Daily Exposure

0

0.2

0.4

0.6

0 5 10 15 20 25 30

Time (day)

|Cor

rela

tion

Coe

ffici

ent|

PearsonSpearman

95% Confidence Interval of the Coefficient

45

Figure 5-4(c) illustrates another typical pattern for inputs that vary on a daily basis. For

an individual day, such inputs can be quite important, as indicated in the result for the first day.

The correlations here are approximately 0.4 for PCC and 0.5 for SCC, which are relatively high.

By the second day, the correlation is only approximately 0.2 for either type of correlation, and by

the 6th day the correlations are statistically insignificant in both cases. A key assumption in the

simulation is that this input has independent values from one day to the next. Thus, it is possible

that the input could have a high value on one day, and a low value on the next day, or vice versa.

When comparing Scenario I and III, it is apparent that this input is important on the first day, and

declines significantly in importance with respect to daily exposures. Simultaneously, the

magnitude of daily exposures typically decreases with time. Thus, the cumulative effect tends to

decrease. However, unlike the situation for inputs that remain constant during the month, an

additional factor is likely to be important.

The additional factor is the effect of averaging over the course of many days for those

inputs that vary on a daily basis. This effect is easiest to identify in a purely linear model. As an

input is sampled independently from one day to the next, its values will fluctuate randomly

among high and low values according to the frequency distribution from which daily values are

generated. However, by the end of the month, the overall effect of such an input will be based

upon the sampling distribution of the mean. The sampling distribution of the mean has a much

smaller variance than does the frequency distribution of the daily variability. Thus, by the end of

the month, the effective range of variation for a given individual tends to narrow and converge

with respect to the mean. Furthermore, if all individuals are assumed to have independent

random samples from the same frequency distribution, as is the case here, then the overall effect

is to arrive at a sampling distribution of the mean based upon 300,000 samples. The standard

46

error of a sampling distribution with such a large number of samples is quite small – in this case,

less than approximately 0.2 percent of the population standard deviation. Thus, daily inputs

sampled in this way are destined to become unimportant with respect to cumulative exposure

unless the model is highly nonlinear or non-independent sampling schemes are used.

For Scenarios I and III, the SCC and PCC correlation coefficients had similar

magnitudes, and their confidence intervals typically overlapped, except on the first day. This is

consistent with observations for the previously discussed inputs that these two types of

correlations produce similar numerically values when correlations are small, but they tend to

differ when correlations are large and presumably when there is a nonlinearity characteristic to

the model. For Scenario II, for which the SCCs tend to be consistently at a value of

approximately 0.5, there is a more pronounced difference compared to the PCCs. In fact, not

only the numerical values but the trends appear to be different. For example, while the SCCs are

approximately constant over all days of the month, the PCCs appear to decrease monotonically

toward values that, while statistically significant, imply a weak correlation. A possible reason

for this difference is an overall change in the behavior of the model later in the month. As noted

earlier in the discussion of residue decay rate, the model becomes more sensitive to the residue

decay rate on a daily basis. The model has a nonlinear response to residue decay rate that

becomes more pronounced later in the month. Thus, there may be an interaction effect

associated with this source of nonlinearity, which could explain why the SCCs and PCCs for

other inputs, such as FTR, diverge more sharply later in the month.

A detailed examination of the temporal profile of sensitivity results for selected inputs

illustrates the complexity of the temporal dynamic of sensitivity, the competition among inputs,

47

and the changing response of the model as different components or aspects (e.g., nonlinearity) of

the model become more dominant.

5.2.2 Rankings for Selected Inputs

At each time step, inputs were ranked ordered based on the relative magnitude of

correlation coefficients. A rank of one is assigned to an input with highest correlation coefficient,

and the largest numerical value of rank was assigned to the input with the least importance.

However, inputs that have ranks closest to one (that are highly important) are typically described

as inputs that are “highly ranked” or that have “higher” or “high” ranks, whereas inputs with

large numerical values of rank (implying lack of importance) are typically described as having

“low” rank. The rank for each input typically changed with respect to time. Some inputs had

higher ranks in earlier time steps of the model simulation, while their sensitivity declined later in

time. In contrast, some other inputs were typically unimportant in earlier time steps, while their

sensitivity improved later in time. An example for each case is provided. Examples provided in

the following are based on those shown in Figures 5-2 and 5-4, for residue decay rate (DR) and

fraction of chemicals available for transfer (FTR).

Figure 5-5 shows the variation in rank of DR with respect to time for three temporal

scenarios. The ranks are based on PCC and SCC in Figures 5-5 (a) and 5-5 (b), respectively. DR

is an example of an input that had less sensitivity compared to other inputs earlier in the month,

while its sensitivity increased later in the month. The most significant findings are that DR

initially was of only moderate importance to both daily and cumulative exposures early in the

month, but became more important later in the month for reasons previously discussed. Late in

the month, DR was typically the top ranked input. Of course, the magnitude of exposure

decreases later in the month, and thus the daily exposures become less of a factor in terms of

48

(a) Pearson Correlation

0

510

1520

25

1 5 10 15 20 25 30

Time (day)

Ran

kScenario I Scenario II Scenario III

(b) Spearman Correlation

0

5

101520

25

1 5 10 15 20 25 30

Time (day)

Ran

k

Scenario I Scenario II Scenario III

Figure 5-5. Variation in Rank of Residue Decay Rate as an Input with Monthly Sampling Strategy with Respect to Time Based on:

(a) Pearson; and (b) Spearman Correlation Coefficients.

Lower Rank Implies Greater Sensitivity

49

influence on the cumulative exposure. Nonetheless, in terms of overall impact on cumulative

exposure, DR is the most important input by Day 6 and at the end of the month. The rankings for

cumulative exposure at the end of the month are likely to be the most indicative of the overall

importance of various inputs with respect to cumulative exposure.

The rank of DR with respect to incremental changes in exposure (Scenario II) has an

interesting pattern. Initially, DR is of moderate or minor importance, with a rank of

approximately 10. By the seventh day, the rank worsens, and DR becomes one of the least

important inputs. By the end of the month, however, DR is one of the most important inputs.

However, the distinction between a rank in the top five versus a rank of approximately 10 is

more pronounced than that between a rank of 10 and one of the lowest ranks. In other words,

there is more difference in sensitivity between a top rank and moderate rank than between a

moderate rank and a low rank. Thus, in approximate terms, DR is not a particularly important

input with respect to incremental change until late in the month. As noted in previous text, the

model becomes more strongly sensitive to DR later in the month, and the model response to DR is

nonlinear. Thus, it appears to be the case that incremental changes in exposure late in the month

are highly influenced by the incremental change in pesticide residue remaining, which in turn is

highly influenced by the residue decay rate.

Figure 5-6 shows the variation in rank of fraction of chemicals available for transfer (FTR)

with respect to time for the three temporal scenarios based on PCC and SCC. FTR is an example

of an input that had relatively higher sensitivity earlier in the month, while its sensitivity declined

later in the month for Scenarios I and III. Furthermore, unlike DR, FTR has a daily sampling

frequency. On the first day, FTR is the most important input for all three temporal scenarios. By

day 2 or 3, however, FTR begins to decline in its rank order, especially with respect to the

50

(a) Pearson Correlation

0

510152025

1 5 10 15 20 25 30

Time (day)

Ran

kScenario I Scenario II Scenario III

(b) Spearman Correlation

0

5

101520

25

1 5 10 15 20 25 30

Time (day)

Ran

k

Scenario I Scenario II Scenario III

Figure 5-6. Variation in Rank of Fraction of Chemical Available for Transfer as an Input with Daily Sampling Strategy with Respect

to Time Based on: (a) Pearson; and (b) Spearman Correlation Coefficients.

Lower Rank Implies Greater Sensitivity

51

cumulative exposure. By approximately the 5th day, FTR reaches a consistent ranking in Scenario

I of approximately the fifth most important input. However, by the 10th day, FTR becomes one of

the least important inputs with respect to cumulative exposure. Although the ranking in Scenario

III appears to be “noisy,” with what appear to be large changes from day-to-day, rankings that

are higher in numerical value than approximately 15 are associated with inputs that have either

minor or insignificant sensitivity. Thus, large changes in rank order among insignificant inputs

does not imply any significant change in importance. In contrast, small changes in rank order

among important inputs can imply a significant change in importance. Furthermore, the fact that

FTR is unimportant on the 30th day for the cumulative exposure suggests that this input has little

overall effect on cumulative exposure. If one is concerned with very short-term acute exposures,

the fact that this input is the most important on the first day, when exposures tend to be high,

would be of significance. However, if one is concerned with average or cumulative exposures

over many days, then this input is not likely to be of much importance.

As noted earlier in the discussion of the magnitude of the correlation coefficients, FTR is

consistently the most important input with respect to incremental change in daily exposures

(Scenario II). This may be useful information for understanding what factors govern the rate at

which daily exposures are changing, and thus understanding the behavior of the model.

However, the incremental changes from one day to the next may not be the dominant cause for

variation in exposures among individuals in a given day nor on the cumulative exposure. For

example, DR has a persistent effect on inter-individual variability in daily and cumulative

exposures even though from one day to the next it does not cause a large change in exposure for

any given individual. Thus, DR would likely be considered to be a more significant or important

input than FTR.

52

5.2.3 Comparison of Mean Ranks

This section presents the mean rank associated with each input within the time period of

the model simulation. The mean rank represents the arithmetic average of ranks for each input.

The range of ranks also is given for each input. The range of ranks represents minimum and

maximum ranks for the input within 30 days of model simulation. However, a rank of 1 on day

one might be much more significant than a rank of 1 on day 30, because the daily exposure on

day one is much higher than on day 30, and the effect on the cumulative exposure is more

pronounced early in the month than later in the month. Despite this limitation the use of a mean

rank and a range of ranks was deemed to be a useful way to provide a semi-quantitative

comparison of the overall significance of inputs during the course of the 30 day simulation

period.

Figure 5-7 (a) shows the mean rank and inter-daily range of ranks for each input based on

PCC. The input abbreviations are given in Table A.1 in Appendix A. On average, the most

sensitive input in Scenario I was probability of washing body (PWB) with a mean rank of 1.8. The

range of ranks for this input was between 1 and 2, and hence, PWB was among the top two

sensitive inputs throughout the entire month. Inputs with secondary sensitivity with respect to

daily exposures were residue decay rate (DR), body weight (WB), amount of pesticide applied

(AM), fraction of chemicals available for transfer (FTR), and the application area (Area). Average

ranks for these inputs within 30 days of model simulation ranged between 2.1 and 4.7. These

inputs typically had wide ranges of ranks.

In Scenario II, FTR was identified as the most sensitive input throughout the entire month,

and hence, no range of ranks was associated with this input. Transfer coefficient via body

(TCBody), exposure duration at target area (EDTarget), and body washing removal efficiency (BW)

53

(a) Pearson Correlation

0

5

10

15

20

25

Cai

r

Cb

WB k

PH

PW

B

AM

Are

a

RN

/T

Hei

ght

DR

PA

I

Din

h

HW

BW LH LB E

f

TCha

nd

TCbo

dy

ED

targ

et

ED

non-

targ

et

FTR

FBU HF

Inputs

Mea

n R

ank

Scenario I Scenario II Scenario III

(b) Spearman Correlation

0

5

10

15

20

25

Cai

r

Cb

WB k

PH

PW

B AM

Are

a

RN

/T

Hei

ght

DR

PA

I

Din

h

HW

BW LH LB E

f

TCha

nd

TCbo

dy

ED

targ

et

ED

non-

targ

et

FTR

FBU HF

Inputs

Mea

n R

ank

Scenario I Scenario II Scenario III

Figure 5-7. Comparison of Mean Ranks and Range of Ranks Representing Minimum and Maximum Ranks of Inputs in 30 Days of Model Simulation for the Three Temporal Scenarios Based on: (a) Pearson Correlation Coefficients; and (b) Spearman Correlation

Coefficients.

54

were identified as inputs with secondary sensitivity with mean ranks between 3.1 and 5.6. Inputs

that were identified with secondary sensitivity in Scenario I, typically were among the inputs

with low sensitivity with respect to the incremental change in daily exposures. For example, DR

had a mean rank of 8.4 in Scenario II with a range of ranks between 2 and 20. However, DR has

a montly sampling frequency and for reasons previously discussed would not be expected to be

of significant importance in Scenario II.

In Scenario III, the inputs that have the most importance and secondary importance are

approximately the same as those for Scenario I y. The only exception was with respect to

sensitivity of FTR to cumulative daily exposure. FTR had a low sensitivity with a mean rank of

14.2 and a range of ranks between 1 and 24 in Scenario III.

However, as noted in the previous section, if the main interest is with respect to which

inputs are important to cumulative exposure, then the rankings for cumulative exposure at the

end of the overall time period of interest (e.g., one month in this case) may be of more usefulness

than the average daily rank.

Figure 5-7 (b) shows the mean rank and corresponding range of ranks for each input

based on SCC. Typically, SCC provided the same results with respect to inputs with the most

and secondary sensitivity in the three temporal scenarios. However, comparison of results for the

two correlation-based methods shows that typically inputs that had high sensitivity based on the

two methods, on average had lower mean ranks (i.e., higher sensitivity) based on SCC. For

example, amount of pesticide applied on a monthly basis (AM) had mean ranks of 3.3, 9.9, and

2.6 in the three temporal scenarios, respectively, based on PCC. However, the mean ranks for AM

were 2.2, 7.5, 1.5 in the three temporal scenarios, respectively, based on SCC. Moreover, the

range of ranks associated with each input was narrower when Spearman correlation coefficients

55

were used to rank the inputs. For example, AM was ranked between 2 and 24 and between 3 and

16 based on PCC and SCC, respectively, in Scenario II. An implication is that the SCC results

may be more “stable” than the PCC results, and thus more robust. Of course, this can be due to

nonlinearity effects in the model, which are better accounted for using SCC than PCC.

5.2.4 Summary of Results Based on Correlation Coefficients

Summary of key insights and findings based on correlation-based methods include:

• The magnitude of correlation coefficients typically changed for inputs with respect to

time, regardless of whether inputs were sampled on a monthly or daily basis. Thus, there

is clearly a temporal dynamic in the case study. Sensitivity analysis provides insight and

understanding regarding the causes of the temporal dynamic, and regarding the relative

importance of inputs at different times of the month and with respect to different methods

of quantifying exposure over time (e.g., Scenarios I, II, an d III).

• For inputs that were sampled monthly, there were some general trends in the temporal

pattern of sensitivity. Monthly inputs can either increase or decrease in sensitivity over

time with respect to daily exposures. However, they tend to be unimportant, except

perhaps in the first few days, with regard to incremental changes in exposures. This

result is not inconsistent with situations in which an input is important on a daily basis.

For example, a monthly input can persistently contribute to inter-individual variability in

daily exposures from one day to the next, but might not contribute significantly to

marginal changes in exposures for a given individual from one day to the next.

• Monthly inputs tend to increase in sensitivity with respect to cumulative exposures during

the first few days of the month. For those monthly inputs that increase in importance for

daily exposure as time elapses, there will be a monotonic increase in sensitivity for

56

cumulative exposure as well. However, the marginal increase in sensitivity in such cases

diminishes, at least in the case studies here, because the daily exposures decrease in

magnitude with time. Therefore, there is a dimishing contribution to the cumulative

exposure.

• Correlation coefficients for inputs with a daily sampling strategy typically declined

monotonically with respect to time for both daily and cumulative exposures. A key

reason for this is because of the independence assumption assumed in the case study,

leading to random fluctuations in daily values of such inputs for each individual. For a

linear model in particular, the fluctuations tend to “average out,” with the net effect being

that the influence of the input is really based on the sampling distribution of its mean,

rather than on variability in individual sample values.

• Although inputs sampled with a daily frequency typically become unimportant in

Scenarios I and III by late in the month, they can be important for Scenario II throughout

the month. This is because such inputs can significantly influence day-to-day changes in

the magnitude of daily exposures, even though they may not significantly influence inter-

individual variability in exposures because of the averaging effect discussed above.

• When comparing SCC to PCC, the rank transformation typically led to correlation

coefficients of equal or greater magnitude. For very small correlations, the two methods

provide approximately the same numerical result. For larger correlations (e.g., above

approximately 0.2), the two methods diverge in results. For correlations with a large

magnitude, the SCC coefficients were often statistically significantly greater than those

from PCC. The numerical values at which the two coefficient estimates diverge may be

in part an artifact of the specific model and input assumptions used here. However, in

57

general, for a model that deviates in some way from pure linearity, but that is monotonic,

the SCC is expected to provide a larger numerical value than the PCC for strong

dependencies.

• An implication of the previous finding is that both PCC and SCC may be equally capable

of identifying inputs that are little or no significance. This could be important, for

example, if these methods are used as a screening step prior to application of a more

refined, but also more computationally intensive, sensitivity analysis method. For

example, if many inputs are found to be insignificant or unimportant, their variation

could either be ignored or they could be set to point estimates, and then more

computationally intensive but more accurate sensitivity analysis methods can be applied

to quantify and discriminate the importance among those inputs of high or secondary

importance.

• The SCC appears to be a more robust measure of sensitivity than the PCC, in that there is

typically less variability in the ranks that are inferred by comparing SCCs for all inputs

on a given day than there is based on the PCCs. Thus, the use of SCC appears to produce

more consistent findings regarding the relative importance of each input. As such, SCC

would typically be preferred over PCC.

In general, the results based on correlation coefficients of both types were approximately

comparable, but preference should be given to SCC over PCC at least with regard to applicability

and robustness for the type of modeling used as a testbed here.

In the following sections, the results of additional methods are introduced. This section

provided detailed interpretation and explanation of the key trends and findings from a time-based

58

sensitivity analysis approach. These details will not be repeated in the following sections.

Instead, focus will be given to any key differences in interpretation or insights.

5.3 Regression Analysis

This section presents the results based upon the use of standardized sample (SRA) and

rank linear regression analyses (RRA) to characterize the sensitivity of the selected model

outputs for temporal Scenarios I, II, and III to each of the inputs of the model. Linear effects of

the inputs listed in Table A.1 in Appendix A are included in the regression model. For simplicity,

no interaction effects between inputs were considered. Selected inputs that are the same as those

presented in Section 5.2 are considered as the focus of discussion here and specific details are

given.

This section begins with a discussion of the diagnostic checks that can be performed on

regression results prior to interpreting the sensitivity results. In particularly, changes in the

coefficient of determination, R2, of the regression model with respect to time are given (Section

5.3.1). Regression coefficients are given and compared for the selected inputs (Section 5.3.2).

Changes in ranking of selected inputs with respect to time are presented (Section 5.3.3). Mean

ranks and associated range of ranks in 30 days of model simulation for all inputs are also

provided and compared (Section 5.3.4). A summary of the key findings based on the two

regression-based methods are discussed (Section 5.3.5).

5.3.1 Coefficient of Determination

Coefficient of determination, R2, is used to judge the adequacy of the regression model

based on the linear effects of inputs at each time step of the model simulation. R2 represents the

amount of output variability explained or accounted for by the regression model. At each time

step, the coefficient of determination, R2, was estimated for regression models based on both

59

sample and rank transformed data values. Evaluation of the R2 values provided a diagnostic

check on the results. Generally, a high R2 value implies that the regression model captured a

substantial amount of output variability, and hence, the sensitivity analysis results based on the

relative magnitude of regression coefficients deemed to be reliable.

Figure 5-8 shows the variation in R2 values with respect to time. R2 values are shown for

each of the three temporal scenarios. Except for Scenario III, the amount of output variability

that was captured by the regression model based on the sample data (i.e. the sample regression

method) decreased with time. However, rank transformation of the data typically improved the

output variability captured by the regression model. For example, approximately 60 percent of

the output variability was captured by the regression model based on the sample data in the first

10 days in Scenario I, while this amount ranged between low values of 35 and 23 percent in the

last 10 days of the model simulation. Rank transforming data substantially improved the R2

values within the latter time period. The R2 values based on the ranked transformed data ranged

between high values of 0.88 and 0.92 within the last 10 days of the model simulation. R2 values

of 0.9 or higher are considered to be excellent, implying that the regression model used as a basis

for estimating sensitivity is capturing the vast majority of variability in the exposure model, and

thus it is not likely that any significant source of variability in the exposure model results would

be unaccounted for. In contrast, an R2 value of 0.3, such as occurs for sample regression in

60

(a) Scenario I

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30

Time (day)

R2 V

alue

Sample Regression Rank Regression

(b) Scenario II

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30

Time (day)

R2 V

alue

Sample Regression Rank Regression

(c) Scenario III

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30

Time (day)

R2 V

alue

Sample Regression Rank Regression

Figure 5-8. Variation in Coefficient of Determination, R2, with Respect to Time based on Sample and Rank Regression Analyses for: (a) Scenario I: Daily Exposure; (b) Scenario II:

Incremental Change in Daily Exposure; and (c) Scenario III: Cumulative Exposure.

61

Scenario I late in the month, would imply that most of the variation in the exposure model is not

accounted for in the sensitivity analysis. Thus, one might have little confidence that the

sensitivity indices, or the resulting rank order of inputs, would be reliable. Of course, in practice,

the two methods may agree for many kinds of models.

As will be shown later in this section, both sample and rank regression actually provide

similar insights regarding sensitivity. However, as a general rule, and particularly if only one of

these two methods will be applied, it would be prudent to give preference to the method that

more robustly produces reliable results.

For Scenario III, typically R2 values based on sampled and rank transformed data did not

change in time. Moreover, rank transforming data did not improve the amount of output

variability captured by the regression model substantially. The similarity of the two results

implies that the cumulative exposure estimate is more linear than the daily exposure estimate.

This is likely because a cumulative exposure is inherently additive, based upon summing the

daily exposures. Thus, even though there maybe nonlinear effects between daily exposure and a

given input, these nonlinearities may tend to “average out” when adding exposures over multiple

days. Furthermore, even though the diagnostic check for the daily exposures implies significant

divergence between sample and rank regression late in the month, the magnitude of daily

exposures late in the month is small. Therefore, the effect of nonlinearities in daily exposure late

in the month has minor or little effect on cumulative exposure during that time.

5.3.2 Regression Coefficients

For each of the three temporal scenarios, regression coefficients were estimated at each

day for all inputs, and hence, 30 sets of standardized sample regression coefficients and rank

regression coefficients were estimated. Typically, regression coefficients based on both sampled

62

and rank transformed data varied in time in the three temporal scenarios. Similar to correlation-

based methods, for inputs with a monthly sampling strategy, two distinct patterns with respect to

the change in the absolute values of regression coefficients were identified for Scenario I: (1)

monotonic increase; and (2) monotonic decrease. In contrast, inputs with a daily sampling

strategy typically had coefficients with decreasing trend with respect to time for Scenario I.

Examples are provided to illustrate these patterns.

Figure 5-9 shows variation in absolute values of regression coefficients based on SRA

and RRA and corresponding 95% confidence intervals for the residue decay rate (DR) with

respect to time. The coefficients are shown for the three temporal scenarios. DR has a monthly

sampling strategy and provides an example in which the input has relatively higher regression

coefficient later in the month. Overall, the trends here, and the explanations for the tends, are the

same as for the correlation coefficients for each of the three scenarios.

Figure 5-9 (a) shows the results for Scenario I. DR has a substantially higher regression

coefficient later in the month. For example, regression coefficients based on SRA and RRA had

low values of 0.04 and 0.06, respectively, on Day 1. However, there was substantial increase in

the magnitude of these coefficients by the end of the month. The regression coefficient for DR

reached high values of 0.4 and 0.9 based on SRA and RRA, respectively, by the end of the

month.

Figure 5-9 (b) shows the results for Scenario II for the change in regression coefficients

of DR based on SRA and RRA. These results are qualitative similar to those for correlation

coefficients, including the trends over time and that the rank-based method provided larger

magnitudes of the sensitivity index later in the month than did the sample-based method.

63

(a) Scenario I: Daily Exposure

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30Time (day)

|Reg

ress

ion

Coe

ffici

ent|

Sample Regression Rank Regression

(b) Scenario II: Incremental Change in Daily Exposure

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30Time (day)

|Reg

ress

ion

Coe

ffici

ent|

Sample Regression Rank Regression

(c) Scenario III: Cumulative Exposure

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30Time (day)

|Reg

ress

ion

Coe

ffici

ent|

Sample Regression Rank Regression

Figure 5-9. Absolute Values and Corresponding 95% Confidence Intervals of Sample and Rank Regression Coefficients for Residue Decay Rate as an Input with a Monthly Sampling Strategy

for: (a) Scenario I: Daily Exposure; (b) Scenario II: Incremental Change in Daily Exposure; and (c) Scenario III: Cumulative Daily Exposure.

95% Confidence Interval of the Coefficient

64

Figure 5-9 (c) shows the results for Scenario III. Absolute values of regression

coefficients for DR are shown for SRA and RRA along with their 95% confidence intervals in 30

days of model simulation. The temporal trend and the comparison between rank and sample

based methods are similar to those for PCC and SCC as explained in the previous section.

Comparing regression coefficients based on SRA and RRA for the three temporal

scenarios indicates that rank transforming typically increased the magnitude of the coefficients

for DR. However, this increase in the magnitude of the regression coefficients is more

pronounced with respect to Scenario I. Considering high values of R2 for the regression model

based on the rank transformed data in Scenario I (Figure 5-8), regression coefficients based on

RRA provides a much more reliable measure of the sensitivity of daily exposures to DR than

based on SRA.

For Scenario II, rank transforming the data did not improve the magnitude of regression

coefficients. Because R2 values for the regression model in this scenario were typically low, it

matters less as to what type of regression analyses is used.

In Scenario III, rank transforming the data had a minimal effect on improving the

magnitude of regression coefficients, similar to its effect on R2 values for this scenario.

Confidence intervals for the regression coefficients provide insight with respect to

statistically significant differences between coefficients. Typically, regression coefficient based

on SRA and RRA were comparable in the first few days of the model simulation due to

overlapping confidence intervals. However, later in the month, typically these two regression-

based methods provided coefficients which were statistically significantly different for all three

Scenarios. However, for Scenario III, the magnitude of the differences were less pronounced,

even though they are statistically significant.

65

Significant changes in regression coefficients from one time period to another can also be

assessed by confidence intervals for each of the regression-based methods. For example, for

Scenario I, regression coefficients based on the rank transformed data were statistically different

within the first 15 days of the model simulation. However, through the rest of the month, these

coefficients were typically comparable for time periods of several adjacent days.

Figure 5-10 shows variation in absolute values of regression coefficients based on SRA

and RRA and corresponding 95% confidence intervals for probability of washing body (PWB)

with respect to time. PWB has a monthly sampling strategy and provides an example in which the

highest regression coefficient for Scenario I occurs early in the month. Overall, the results and

key findings are approximately similar to those obtained from PCC and SCC, including the

temporal trends in each scenario and the comparison of rank versus sample-based methods.

In Scenario I, the magnitude of regression coefficients is relatively modest, ranging from

approximately 0.1 to 0.2. However, the coefficients are significantly larger within the first 10

days when compared to later in the month. The reason for the peak in the sensitive indices near

the 5th day is explained in the section on PCC and SCC. For Scenario II, the regression

coefficients decline from the first day to about the 5th to 10th day, and there after remain

approximately constant at values that are either practically or statistically insignificant.

The magnitude of regression coefficients increases with time with respect to cumulative

exposure in Scenario III, as shown in Figure 5-10 (c), during the first five days of the month.

However, through the rest of the month, there is no significant change in regression coefficients

based on either method, for reasons explained previously.

For the two regression-based methods, typically regression coefficients were

comparable in different time steps of the model simulation. Confidence intervals for the

66

(a) Scenario I: Daily Exposure

0

0.1

0.2

0.3

0 5 10 15 20 25 30Time (day)

|Reg

ress

ion

Coe

ffici

ent|

Sample Regression Rank Regression

(b) Scenario II: Incremental Change in Daily Expsoure

0

0.1

0.2

0.3

0 5 10 15 20 25 30Time (day)

|Reg

ress

ion

Coe

ffici

ent|

Sample Regression Rank Regression

(c) Scenario III: Cumulative Exposure

0

0.1

0.2

0.3

0 5 10 15 20 25 30Time (day)

|Reg

ress

ion

Coe

ffici

ent|

Sample Regression Rank Regression

Figure 5-10. Absolute Values and Corresponding 95% Confidence Intervals of Sample and Rank Regression Coefficients for Probability of Washing Body as an Input with a Monthly

Sampling Strategy for: (a) Scenario I: Daily Exposure; (b) Scenario II: Incremental Change in Daily Exposure; and (c) Scenario III: Cumulative Daily Exposure.

95% Confidence Interval of the Coefficient

67

regression coefficients based on the two methods typically overlapped. In general, the regression

coefficients were relatively small for all scenarios and time steps. Thus, it appears to be the case

that for an input that has a small sensitivity index, the two methods provide comparable results.

Typically, the regression coefficients based on SRA and RRA did not statistically differ

from one time period to another in the second half of the month for the three temporal scenarios.

Compared to DR, the main difference is that PWB decreases in sensitivity on a daily basis, where

as DR increased in sensitivity on a daily basis, as explained in more detail in Section 5.2.

Figure 5-11 illustrates an example of change in regression coefficients for the fraction of

chemicals available for transfer (FTR). This input has a daily sampling strategy. When compared

to the results from correlation coefficients as shown in Figure 5-4, the general trends are

approximately the same. However, the regression results appear to be more consistent, with less

noisy perturbations, and the sensitivity measures appear to have narrower confidence intervals on

a relative basis. Thus, the sensitivity indices based on regression coefficients appear to be

perhaps more stable and more robust than those based on correlation coefficients.

Except for the first few days in Scenarios I and III, regression coefficients based on SRA

and RRA were statistically comparable. In Scenario II, rank transforming the data significantly

increased the magnitude of the regression coefficients. This is similar to the findings in Section

5.2.

5.3.3 Rankings for Selected Inputs

At each time step, inputs were ranked based on the relative magnitude of the absolute

regression coefficients. The rank of one is assigned to an input with highest regression

coefficient. A similar pattern compared to rankings based on correlation-based methods is

observed for the results from SRA and RRA. Some inputs had relatively higher sensitivity in the

68

(a) Scenario I: Daily Exposure

0

0.2

0.4

0.6

0 5 10 15 20 25 30Time (day)

|Reg

ress

ion

Coe

ffici

ent|

Sample Regression Rank Regression

(b) Scenario II: Incremental Change in Daily Exposure

0

0.2

0.4

0.6

0 5 10 15 20 25 30Time (day)

|Reg

ress

ion

Coe

ffici

ent|

Sample Regression Rank Regression

(c) Scenario III: Cumulative Exposure

0

0.2

0.4

0.6

0 5 10 15 20 25 30Time (day)

|Reg

ress

ion

Coe

ffici

ent|

Sample Regression Rank Regression

Figure 5-11. Absolute Values and Corresponding 95% Confidence Intervals of Sample and

Rank Regression Coefficients for Fraction of Chemicals Available for Transfer as an Input with a Daily Sampling Strategy for: (a) Scenario I: Daily Exposure; (b) Scenario II: Incremental

Change in Daily Exposure; and (c) Scenario III: Cumulative Daily Exposure.

95% Confidence Interval of the Coefficient

69

(a) Sample Regression

0

5

1015

20

25

1 5 10 15 20 25 30

Time (day)

Ran

kScenario I Scenario II Scenario III

(b) Rank Regression

05

10

15

20

25

1 5 10 15 20 25 30

Time (day)

Ran

k

Scenario I Scenario II Scenario III

Figure 5-12. Variation in Rank of Residue Decay Rate as an Input with Monthly Sampling Strategy with Respect to Time Based on: (a) Sample Regression Analysis; and (b) Rank

Regression Analysis.

earlier time steps, while other inputs showed higher sensitivity in later time steps of the model

simulation. Typically, inputs with monthly sampling frequencies had higher sensitivity in later

time steps, while inputs with daily sampling frequencies were among the list of inputs with top

sensitivity in earlier time steps of the model simulation. Residue decay rate (DR) and the fraction

of chemicals available for transfer (FTR) are selected as two examples in order to illustrate the

pattern of change in sensitivity of an input on a daily basis.

Figures 5-12 shows variation in ranks of DR with respect to time for the three temporal

scenarios. The ranks are based on SRA and RRA in Figures 5-12 (a) and 5-12 (b), respectively.

DR typically showed lower sensitivity earlier in the month, while its sensitivity increased later in

the month for all three temporal scenarios. For example, for Scenario II, while DR had a range of

Lower Rank Implies Greater Sensitivity

70

(a) Sample Regression

05

10

15

20

25

1 5 10 15 20 25 30

Time (day)

Ran

kScenario I Scenario II Scenario III

(b) Rank Regression

0

5

10

15

20

25

1 5 10 15 20 25 30

Time (day)

Ran

k

Scenario I Scenario II Scenario III

Figure 5-13. Variation in Rank of Fraction of Chemical Available for Transfer as an Input with

a Daily Sampling Strategy with Respect to Time Based on: (a) Sample Regression Analysis; and (b) Rank Regression Analysis.

ranks between 5 and 24 in the first 16 days of the model simulation based on SRA, it was

identified as the second most sensitive input through out the rest of the month.

Figure 5-13 shows an example in which selected input had relatively higher sensitivity in the

earlier time steps of the model simulation. FTR has a daily sampling frequency and typically had

higher sensitivity within the first 5 days of the model simulation compared to the rest of the

month for daily and cumulative exposure scenarios. However, FTR was identified as the most

important input throughout the entire month in Scenario II. Although the qualitative trends for

the rankings over time are similar to those obtained from the correlation methods, the

quantitative results appear to have some significant differences. For example, in Figure 5-6, the

rankings for Scenario III imply little or no importance in the second half of the month. However,

in Figure 5-13, the rankings for Scenario III imply moderate importance, with ranks typically of

approximately seven. Furthermore, there is less inter-daily variability in the rankings for

Lower Rank Implies Greater Sensitivity

71

Scenario III in the second half of the month. This implies that the regression method is

responding differently than the correlation methods. However, the R2 values for Scenario III are

typically around 0.6 to 0.8. This implies that the regression model captures or explains most of

the variability in the exposure model, but that a substantial portion of variability is not addressed.

Thus, while the results from the regression methods may appear to be more stable, it is not clear

that they are any more reliable than the correlation methods. In either case, the sample-based

methods presume a linear response and the rank-based methods presume a monotonic response.

Thus, the major theoretical underpinnings with respect to sensitivity analysis are approximately

the same

5.2.4 Comparison of Mean Ranks

This section presents the mean rank associated with each input within the time period of the

model simulation. The mean rank represents the arithmetic average of ranks for each input. The

range of ranks also is given for each input. The range of ranks represents minimum and

maximum ranks for the input within 30 days of model simulation. In general qualitative terms,

the results for the mean ranks shown in Figure 5-14 based on the regression methods are

comparable to those shown in Figure 5-7 for the correlation methods. However, there are some

quantitative differences.

Figure 5-14 (a) shows the mean rank and corresponding range of ranks for each

input based on SRA. The input abbreviations are given in Table A.1 in Appendix A. On average,

the most sensitive input in Scenario I was fraction of chemicals for transfer (FTR) with a mean

rank of 2.0. The range of ranks for this input was between 1 and 4. Inputs with secondary

sensitivity with respect to daily exposures were residue decay rate (DR), amount of pesticides

applied (AM), area of pesticide application (Area), body weight (WB), and probability of

72

(a) Sample Regression Analysis

0

5

10

15

20

25

Cai

r

Cb

WB k

PH

PW

B

AM

Are

a

RN

/T

Hei

ght

DR

PA

I

Din

h

HW

BW LH LB E

f

TCha

nd

TCbo

dy

ED

targ

et

ED

non-

targ

et

FTR

FBU HF

Input

Mea

n R

ank

Scenario I Scenario II Scenario III

(b) Rank Regression Analysis

0

5

10

15

20

25

Cai

r

Cb

WB k

PH

PWB

AM

Are

a

RN

/T

Hei

ght

DR

PA

I

Din

h

HW

BW LH LB E

f

TCha

nd

TCbo

dy

ED

targ

et

EDno

n-ta

rget

FTR

FBU HF

Input

Mea

n R

ank

Scenario I Scenario II Scenario III

Figure 5-14. Comparison of Mean Ranks and Range of Ranks Representing Minimum and Maximum Ranks of Inputs in 30 Days of Model Simulation for the Three Temporal Scenarios Based on: (a) Sample Regression Analysis; and (b) Rank Regression Analysis.

73

washing body (PWB). Average ranks for these inputs within 30 days of model simulation ranged

between 2.3 and 5.5. These inputs typically had wide ranges of ranks.

In Scenario II, FTR was identified as the most sensitive input throughout the entire month,

and hence, no range of ranks was associated with this input. Transfer coefficient via body

(TCBody), exposure duration at target area (EDTarget), and body washing removal efficiency (BW)

were identified as inputs with secondary sensitivity with mean ranks between 4.7 and 5.4. Inputs

that were identified with secondary sensitivity in Scenario I typically were among the inputs with

low sensitivity with respect to the incremental change in daily exposures. For example, DR had a

mean rank of 8.0 in Scenario II with a range of ranks between 2 and 24.In Scenario III, AM was

identified as the most sensitive input with a mean rank of 2.0 in 30 days of model simulation.

Inputs with secondary sensitivity included WB, DR, Area, and FTR with mean ranks between 2.7

and 5.2.

Figure 5-14 (b) shows the mean rank and corresponding range of ranks for each input

based on RRA. Typically, RRA provided the same results with respect to inputs with the most

and secondary sensitivity in the three temporal scenarios. However, comparison of results for the

two regression-based methods shows that typically inputs that had high sensitivity based on the

two methods, on average had lower mean ranks (i.e., higher sensitivity) based on RRA.

Moreover, the range of ranks associated with each input was narrower when rank regression

coefficients were used to rank the inputs. This is similar to the findings based on comparison of

SCC to PCC.

When comparing the regression-based results to the correlation-based results, there are

some perhaps minor differences in the rankings. For example, PBW tends to have less sensitivity

according to the regression methods, whereas FTR tends to have slightly more sensitivity. For

74

FTR, all of the methods imply more or less the same overall level of importance. However, for

PBW, the regression methods imply a significantly lower ranking especially for Scenarios I and

III. Because of the very high R2 for RRA in Scenario I, the findings based on this method might

be treated with some deference. However, there does not appear to be much different in average

ranking even for this input when comparing SRA to RRA, despite the much lower R2 for the

former.

5.2.5 Summary of Results Based on Regression-Based Methods

The key insights and findings based on regression-based methods include:

• Similar to the findings with PCC and SCC, the magnitude of regression coefficients

typically changed for inputs with respect to time, regardless of whether inputs were

sampled on a monthly or daily basis. Thus, there is clearly a temporal dynamic in the

case study.

• For inputs that were sampled monthly, there were some general trends in the temporal

pattern of sensitivity. Monthly inputs can either increase or decrease in sensitivity

over time with respect to daily exposures. However, they tend to be unimportant,

except perhaps in the first few days, with regard to incremental changes in exposures.

• Monthly inputs tend to increase in sensitivity with respect to cumulative exposures

during the first few days of the month. For those monthly inputs that increase in

importance for daily exposure as time elapses, there will be a monotonic increase in

sensitivity for cumulative exposure as well. However, the marginal increase in

sensitivity in such cases diminishes, at least in the case studies here, because the daily

exposures decrease in magnitude with time. Therefore, there is a diminishing

contribution to the cumulative exposure.

75

• Regression coefficients for inputs with a daily sampling strategy typically declined

monotonically with respect to time for both daily and cumulative exposures. A key

reason for this is because of the independence assumption assumed in the case study,

leading to random fluctuations in daily values of such inputs for each individual. For

a linear model in particular, the fluctuations tend to “average out,” with the net effect

being that the influence of the input is really based on the sampling distribution of its

mean, rather than on variability in individual sample values.

• Although inputs sampled with a daily frequency typically become unimportant in

Scenarios I and III by late in the month, they can be important for Scenario II

throughout the month. This is because such inputs can significantly influence day-to-

day changes in the magnitude of daily exposures, even though they may not

significantly influence inter-individual variability in exposures because of the

averaging effect discussed above.

• When comparing RRA to SRA, the rank transformation typically led to regression

coefficients of equal or greater magnitude. However, in general, for a model that

deviates in some way from pure linearity, but that is monotonic, the RRA method is

expected to provide a more robust numerical value than the SRA method.

• An implication of the previous finding is that both SRA and RRA may be equally

capable of identifying inputs that are little or no significance. This could be

important, for example, if these methods are used as a screening step prior to

application of a more refined, but also more computationally intensive, sensitivity

analysis method. For example, if many inputs are found to be insignificant or

unimportant, their variation could either be ignored or they could be set to point

76

estimates, and then more computationally intensive but more accurate sensitivity

analysis methods can be applied to quantify and discriminate the importance among

those inputs of high or secondary importance.

In general, the results and findings from the regression methods are similar to those from the

correlation methods. However, there are some differences in quantitative results.

5.4 Analysis of Variance

This section presents the results based upon the use of ANOVA to characterize the

sensitivity of the selected model outputs for temporal Scenarios I, II, and III to each of the inputs

of the model. The main (direct) effects of the inputs listed in Table A.1 in Appendix A are

included in the ANOVA analysis. For simplicity, no interaction effects between inputs were

considered. However, such effects could be considered by expanding the scope of application of

ANOVA to explicitly include interaction terms between selected combinations of inputs.

Inputs in the simplified SHEDS-Pesticides model were continuous, and hence, must be

partitioned into levels prior to application of ANOVA (Kleijnen and Helton, 1999). Frey et al.

(2004) demonstrated three approaches for defining levels for continuous inputs based on: (1)

evenly spaced intervals; (2) evenly spaced percentiles; and (3) visual inspection of the

cumulative distribution function (CDF) for each input. For simplicity and in order to be

consistent with respect to defining levels at different time steps of the model simulation, the

evenly spaced percentile approach was used to define levels. In this approach the CDF of the

generated values for an input in a probabilistic simulation is used for defining levels at evenly

spaced percentiles. For each input, four levels were defined based on the 25th, 50th, and 75th

percentiles of the generated samples at each time step.

77

F values are used to measure sensitivity and are given and compared for the selected

inputs (Section 5.4.1). Changes in ranking of selected inputs with respect to time are presented

(Section 5.4.2). Mean ranks and associated range of ranks in 30 days of model simulation for all

inputs are also provided and compared (Section 5.4.3). A summary of the key findings based on

ANOVA are discussed (Section 5.4.4).

5.4.1 F Values

For each of the three temporal scenarios, F values associated with inputs in the ANOVA

model were estimated at each day, and hence, 30 sets of F values were estimated. Typically, F

values varied in time in the three temporal scenarios. The F values had similar patterns with

respect to change in values in time compared to correlation and regression-based methods

discussed in previous sections. For the purpose of illustration, two examples are given.

Figure 5-15 (a) shows variation in F values of residue decay rate (DR). Similar to the

pattern shown for correlation and regression coefficients, F values for DR typically

monotonically increased in time for the three temporal scenarios. However, the magnitude of F

values was substantially higher in Scenario I compared to the other two scenarios. F Values for

DR reached a high value of approximately 350 on Day 10 and maintained approximately the

same high value through the rest of the month. The trends for Scenarios II and III imply the

same general findings as were obtained from either correlation or regression methods.

Figure 5-15 (b) shows variation in F values for fraction of chemicals available for transfer

(FTR) with respect to time. In contrast to DR, F values corresponding to FTR were larger in earlier

time steps of the model simulation, and declined later in the month similar to the pattern shown

for correlation and regression coefficients in Sections 5.2 and 5.3. Because rankings are based on

the relative magnitude of F values at each time step, these inputs are expected to have different

78

(a) Residue Decay Rate

0

50

100

150

200

250

300

350

400

0 5 10 15 20 25 30

Time (day)

F Va

lue

Scenario I Scenario II Scenario III

(b) Fraction of Chemical Available for Transfer

0

50

100

150

200

250

300

350

0 5 10 15 20 25 30

Time (day)

F Va

lue

Scenario I Scenario II Scenario III

Figure 5-15. Variation of F Values with Respect to Time for the Three Temporal Scenarios for: (a) Residue Decay Rate; and (b) Fraction of Chemicals Available for Transfer.

sensitivity throughout the month. The rankings of these two inputs are presented in the next

section.

5.4.2 Rankings for Selected Inputs

At each time step, inputs were ranked based on the relative magnitude of the F values. A

rank of one is assigned to an input with highest F value. The change in ranks for input with

respect to time followed a similar pattern compared to correlation and regression-based methods.

79

(a) Residue Decay Rate

05

10152025

1 5 10 15 20 25 30

Time (day)

Ran

k

Scenario I Scenario II Scenario III

(b) Fraction of Chemicals Available for Transfer

0

5

10

15

20

25

1 5 10 15 20 25 30

Time (day)

Ran

k

Scenario I Scenario II Scenario III

Figure 5-16. Variation of Ranks with Respect to Time for the Three Temporal Scenarios for: (a)

Residue Decay Rate; and (b) Fraction of Chemicals Available for Transfer.

Some inputs had relatively higher sensitivity in the earlier time steps, while other inputs

showed higher sensitivity in later time steps of the model simulation. Typically, inputs with

monthly sampling frequencies had higher sensitivity in later time steps, while inputs with daily

sampling frequencies were among the list of inputs with top sensitivity in earlier time steps of the

model simulation, particularly for Scenarios I and III. Inputs that were sampled on a daily basis

typically had more influence on the results of Scenario II, particularly later in the month.

Figure 5-16 (a) shows variation in ranks of DR with respect to time for the three temporal

scenarios. DR typically showed lower sensitivity earlier in the month, while its sensitivity

increased later in the month for the three temporal scenarios. For example, while DR had a range

of ranks between 3 and 22 in the first 15 days of the model simulation, it was typically identified

Lower Rank Implies Greater Sensitivity

80

as the second most sensitive input through out the rest of the month in Scenario II. The results

here are very similar to those from correlation and regression methods for all three scenarios.

Figure 5-16 (b) shows an example in which the selected input had relatively higher

sensitivity in the earlier time steps of the model simulation. FTR has a daily sampling strategy and

typically had higher sensitivity within the first 5 days of the model simulation compared

to the rest of the month for daily and cumulative exposure scenarios. However, FTR was

identified as the most sensitive input throughout the entire month in Scenario II. The results

from ANOVA in Figure 5-16 for FTR are quantitatively more similar to those from the

correlation methods in Figure 5-6 than for the regression methods as shown in Figure 5-13. In

particular, the rankings for Scenario III late in the month are very similar for ANOVA and the

correlation methods, with what appears to be more fluctuation in ranks on a daily basis. As

discussed in Section 5.2, these rankings simply imply that there is a group of inputs all of which

are either of minor or insignificant importance, and thus changes in rankings within such a group

of inputs is not of any practical significance.

5.4.3 Comparison of Mean Ranks

This section presents the mean rank associated with each input within the time period of

the model simulation. The mean rank represents the arithmetic average of ranks for each input.

The range of ranks also is given for each input. The range of ranks represents minimum and

maximum ranks for the input within 30 days of model simulation.

Figure 5-17 shows the mean rank and corresponding range of ranks for each input based

on ANOVA. The input abbreviations are given in Table A.1 in Appendix A. On average, the

most sensitive input in Scenario I was residue decay rate (DR) with a mean rank of 1.9. The range

of ranks for this input was between 1 and 10. Inputs with secondary sensitivity with respect to

81

0.0

5.0

10.0

15.0

20.0

25.0

Cai

r

Cb

WB k

PH

PWB

AM

Are

a

RN

/T

Hei

ght

DR

PAI

Din

h

HW

BW LH LB Ef

TCha

nd

TCbo

dy

EDta

rget

ED

non-

targ

et

FTR

FBU HF

Inputs

Mea

n R

anks

Scenario I Scenario II Scenario III

Figure 5-17. Comparison of Mean Ranks and Range of Ranks of Inputs in 30 Days for the Three

Temporal Scenarios Based on ANOVA.

daily exposures were probability of washing body (PWB), amount of pesticides applied (AM),

area of pesticide application (Area), fraction of chemicals available for transfer (FTR), and body

weight (WB). Average ranks for these inputs within 30 days of model simulation were between

2.1 and 5.7.These inputs typically had wide ranges of ranks.

In Scenario II, FTR was identified as the most sensitive input throughout the entire month,

and hence, no range of ranks was associated with this input. Transfer coefficient via body

(TCBody), exposure duration at target area (EDTarget), and body washing removal efficiency (BW)

were identified as inputs with secondary sensitivity with mean ranks between 4.3 and 5.4. Inputs

that were identified with secondary sensitivity in Scenario I, typically were among the inputs

with low sensitivity with respect to the incremental change in daily exposures. For example, DR

had a mean rank of 6.5 in Scenario II with a range of ranks between 2 and 22.

In Scenario III, PWB was identified as the most sensitive input with a mean rank of 1.5 in

30 days of model simulation. Inputs with secondary sensitivity included AM, DR, Area, and WB

with mean ranks between 2.6 and 4.8.

Overall, the average rankings obtained from ANOVA are similar to those for the

correlation methods. The rankings from ANOVA differ quantitatively in some ways from those

82

obtained from the regression methods, in the same manner in which the correlation-based results

differ somewhat from the regression-based results. This implies that ANOVA is providing

insights perhaps more similar to those from the correlation-based than the regression-based

methods. However, in a general sense, all three sets of methods are providing comparable

insights.

Although not explored here quantitatively, ANOVA and regression have the advantage of

being expanded to deal with selected user-specified interaction effects. In this regard, ANOVA

would be expected to have an advantage over regression methods in that a functional form for

the interaction effect need not be assumed.

5.4.5 Summary of Results Based on ANOVA

The key insights and findings from ANOVA are essentially the same as those obtained

from the correlation methods as detailed in Section 5.2.4. The details are not repeated here.

However, the general findings are briefly summarized:

• The magnitude of F values typically changed for inputs with respect to time.

• F values for inputs with a monthly sampling strategy typically in Scenario I: (1) declined

monotonically; or (2) increased monotonically with respect to time.

• F values for inputs with a daily sampling strategy typically declined monotonically with

respect to time in Scenarios I and III, but often are high for Scenario II.

• Results for Scenario II for the top sensitive inputs were typically different from those in

the other two temporal scenarios.

• ANOVA produces results that are more similar to correlation than regression methods.

However, ANOVA has the advantage of additional flexibility compared to correlation

methods, and of being model-independent when compared to regression methods, in that

83

it is possible to add interaction terms and account for these in determining key

associations with the variation in the model output.

5.5 Sobol’s Method

This section presents results based upon the use of Sobol’s method to characterize the

sensitivity of the selected model outputs for temporal Scenarios I, II, and III to each of the inputs

of the model. Unlike the previous methods, Sobol’s method is intended to provide an

apportionment of the variance in the output with respect to model inputs and interaction effects

among model inputs. In order to estimate the uncertainty in main effect and total effect

sensitivity indices, bootstrap simulation with sample size of 500 is performed. An uncertainty

range of ± one standard error around the mean value is reported for each sensitivity index.

Selected inputs the same as those presented in previous sections are used as the focus of

discussion here and specific details are given.

A discussion is given regarding the contribution of inputs to the output variance (Section

5.5.1). Main and total effect sensitivity indices for selected inputs are given and compared

(Section 5.5.2). Changes in ranking of selected inputs with respect to time are presented (Section

5.5.3). Mean ranks and associated range of ranks in 30 days of model simulation for all inputs

are also provided and compared (Section 5.5.4). The key findings based on Sobol’s method are

discussed (Section 5.3.5).

5.5.1 General Insight Regarding Contribution of Inputs to the Output Variance

For each of the three temporal scenarios, Sobol’s method was applied to the model at

each time step. For each time step, bootstrap simulation was performed to quantify uncertainty

ranges of sensitivity indices including main and total effects. Thus, at each time step 500 main

effect and total effect sensitivity indices were estimated for each input. The uncertainties are

reported in terms of plus or minus one standard error. These are not 95 percent confidence

84

intervals. The reason for reporting in this way is that the bootstrap-p method was used. This

method is considered adequate for estimating standard errors but is not as accurate at estimating

confidence intervals. Either a larger number of bootstrap samples and/or an alternative boostrap

method would be needed for that purpose. However, the range of plus or minus one standard

error is illustrative of the range of uncertainty in the estimates.

The main effect of each input represents the fractional contribution of the input to the

output variance. For a linear model with additive terms, the sum of main effects for all inputs

should equal to one. However, in the case of non-linearity and interactions between inputs, the

sum of main effects will be less than one. Figure 5-18 shows the sum of main effects and

corresponding uncertainty ranges at each time step of the model simulation for each of the three

exposure scenarios.

For Scenario I, as shown in Figure 5-18(a), on average, approximately 60 percent of the

output variance was apportioned to the main effects of inputs within the first 10 days of the

model simulation. However, in the last 10 days of the model simulation, the sum of main effects

relatively decreased to values between 30 and 40 percent. This implies that the model is

increasingly nonlinear later in the monthly simulation period. As discussed in Section 5.2

regarding correlation methods, and as reinforced based on results obtained with regression and

85

(a) Scenario I: Daily Exposure

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30Time (day)

Ave

rage

Sum

of M

ain

Effe

cts

(b) Scenario II: Incremental Change in Daily Exposure

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30Time (day)

Ave

rage

Sum

of M

ain

Effe

cts

(c) Scenario III: Cumulative Exposure

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30Time (day)

Ave

rage

Sum

of M

ain

Effe

cts

Figure 5-18. Variation in Average Sum of Main Effects of Inputs in 500 Bootstrap Simulations and Corresponding Uncertainty Ranges with Respect to Time for the Three Temporal Scenarios.

(Intervals shown represent plus or minus one standard error of the mean value).

86

ANOVA methods, the effect of the residue decay rate becomes more pronounced later in the

month, leading to more nonlinearity in the daily exposure estimates. Thus, it is not surprising

that a declining portion of the variance in the model output for Scenario I is captured by linear

effects. Furthermore, the comparison of rank versus sample-based methods implies that the daily

exposure has a strong monotonic response, but is nonlinear.

Figure 5-18 (b) shows the sum of main effects of inputs at each time step in Scenario II.

In Scenario II, contribution of main effects of inputs to the output variance was substantially

lower in comparison to Scenario I. For example, approximately between 10 and 20 percent of the

output variance was apportioned to the main effects of inputs in the last 10 days of the model

simulation. This finding is qualitatively similar to the insight obtained from analysis of R2

values as shown in Figure 5-8. The model appears to have a significant nonlinear response with

respect to incremental changes from one day to the next, perhaps due to significant interactions

among inputs.

Figure 5-18 (c) shows sum of main effects of inputs at each time step in Scenario III. In

Scenario III, the contribution of main effects of inputs to the output variance did not typically

change at different time steps of the model simulation and was approximately 60 percent. These

findings are consistent with the results shown in Figure 5-8 for the variation of R2 values with

respect to time for a regression model that included linear effects of the inputs.

Typically, at each time step, a few inputs had significantly higher contribution to the

output variance compared to the rest of the inputs. To illustrate this, results are shown in Figure

5-19 for Scenario I for three selected days: Day 1, Day 15, and Day 30. The main effect of the

five inputs that had the largest contribution to variance are shown. Also shown is the sum of the

main effects of all other inputs, and the sum of the interaction effects among all of the inputs.

87

(a) Day 1

BW, 8%

Main Effect of Others,

30%

Interactions, 24%

AM, 6%

WB, 6%

DR, 1%

FTR, 25%

(b) Day 15

Main Effect of Others,

24%

Interactions, 30%

DR, 33%

FTR, 7%

BW, 1%

WB, 2%AM, 3%

(c) Day 30

Interactions, 53%

FTR, 1%

DR, 40%

Main Effect of Others,

3% AM, 1%WB, 1%

BW, 1%

Figure 5-19. Contribution Selected Inputs to the Output Variance in Scenario I for Selected Days: (a) Day 1; (b) Day 15; and (c) Day 30.

FTR: Fraction of Chemicals Available for Transfer DR: Residue Decay Rate WB: Body Weight BW: Body Washing Efficiency AM: Amount of Chemical Applied

88

The inputs that typically were in the top five in terms of greatest contribution to the main

effect include fraction of chemicals available for transfer (FTR), residue decay rate (DR), body

weight (WB), body washing efficiency (BW), and amount of chemical applied (AM). Of course,

because the rankings of these inputs change on a daily basis, on some days some of these inputs

may be outside of the top 5.

On Day 1, FTR accounts for 25 percent of the variance in the model output, which is

nearly three times greater than the next most important input (BW). Thus, there is one input that

is clearly a dominate source of variance in comparison to any other individual input. There are

three inputs of comparable but secondary importance, including BW, WB, and AM, which each

account for 6 to 8 percent of the variance of the output, and combined they account for 20

percent of the variance. The fifth input shown individually, DR, is of tertiary importance,

contributing only one percent to the total variance. The other 20 inputs, combined, contribute

only 30 percent to the variance of the output. However, there are significant interaction effects

among all of the inputs. The interaction effects account for 24 percent of the variance in the

output. If there was a need, it is possible to use Sobol’s method to identify the various

interaction effects more specifically, with a trade-off of increased computational effort. For

example, all of the pairwise interactions could be quantified and compared, as could all of the

three-way interactions, four-way interactions, and so on.

FTR contributes to the output variance not only with its main effect, but also through

interactions with other inputs. The sum of the overall contribution of FTR to the output variance is

given in the total sensitivity index of the input. For example, on Day 1 the total effect of FTR is

0.34. This value is given in Figure 5-21 and discussed further later in Section 5.5.2. Thus, on

Day 1, 34% of the output variance was apportioned to the total effect of FTR. Because the total

89

effect of FTR is the sum of main effect and all possible interaction effects involving FTR, on Day

1, FTR has 9% contribution to the output variance via its interaction with other inputs.

On Day 15, the importance of FTR has waned and DR has emerged as the dominant source

of variability in daily exposure. These two inputs account for 40 percent of the variance. The

other three inputs identified specifically account for only 6 percent of the variance, and thus are

of less importance as a group than FTR is on an individual basis. The other 25 inputs contribute

24 percent of the variance, which is less than on Day 1, but the contribution of interactions has

increased to 30 percent. Thus, there is a clear temporal shift among the inputs and with regard to

interaction effects.

By Day 30, DR clearly is the only individual input that can be clearly distinguished from

any other. DR contributes 40 percent to the variance. The next most important input contributes

only 1 percent to the variance. The 20 inputs not explicitly identified contribute only 3 percent

to the variance. However, the interaction effect has become the single most important source of

variance in the output, accounting for just over half of the total variance. As discussed in earlier

sections, the model behavior is more highly nonlinear at this point in time, and thus it is not

surprising that the interaction effect is so pronounced.

The examples shown in Figure 5-19 illustrate the key benefit of Sobol’s method, which is

to provide a relatively clear and intuitive measure of sensitivity. However, when the interaction

effect becomes large, then additional information may be desirable, which could be obtained at

the cost of a much greater computational burden.

5.5.2 Sensitivity Indices

For each of the three temporal scenarios, main effect and total effect sensitivity indices

based were estimated at each day for all inputs. The mean sensitivity index and standard error

90

associated with each index were estimated based on the results of bootstrap at each day. Thus, 30

sets of mean main effects and mean total effects were estimated at each time step of the model

simulation. Typically, sensitivity indices based on the Sobol’s method varied in time in the three

temporal scenarios. The apportionment to variance shown in Figure 5-19 was obtained by

dividing the individual main effect of each input by the sum of the total effects of all inputs.

This section will focus on use of the raw sensitivity index in order to compare total and main

effects for individual inputs and to evaluate temporal trends.

Figure 5-20 shows variation in average main and total effects and corresponding

uncertainty ranges (based on plus or minus one standard error) with respect to time for the

residue decay rate (DR), which has a monthly sampling frequency. In general terms, the results

are qualitatively similar to those from correlation, regression, and ANOVA methods. For

example, as identified based on results from other methods, as time elapses DR becomes more

important later in the month with respect to variability in daily exposures, the model response

becomes more nonlinear, and the magnitude of daily exposures declines. As shown in Figure 5-

20(a), the main effect increases especially in the first 15 days, and then appears to increase only

slowly, if at all, later in the month. However, the total effect, which includes interactions (e.g.,

nonlinearities) continues to increase substantially over the course of the entire month, which

implies increasing nonlinearity that can be attributed to the nonlinear decay term in the model.

The comparison of the main and total effect from Sobol’s method in relation to the rank

versus sample results for correlation and regression methods suggests that the latter may be

offering insight somewhat similar to Sobol’s method. For example, the rank-based methods

would better account for nonlinearities, whereas the sample-based methods focus on the linear

effect. Thus, the rank-based results tend to produce higher coefficients perhaps because they are

91

(a) Scenario I: Daily Exposure

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30

Time (day)

Sen

sitiv

ity In

dex

Main EffectTotal Effect

(b) Scenario II: Incremental Change in Daily Exposure

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30

Time (day)

Sens

itivi

ty In

dex

Main EffectTotal Effect

(c) Scenario III: Cumulative Exposure

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30

Time (day)

Sen

sitiv

ity In

dex

Main EffectTotal Effect

Figure 5-20. Average Main and Total Effects and Corresponding Uncertainty Ranges Based on

500 Bootstrap Simulations for Residue Decay Rate for: (a) Scenario I-Daily Exposure; (b) Scenario II-Incremental Change in Daily Exposure; and (c) Scenario III-Cumulative Exposure.

92

(a) Scenario I: Daily Exposure

0.00

0.20

0.40

0.60

0.80

1.00

0 5 10 15 20 25 30

Time (day)

Sen

sitiv

ity In

dex

Main EffectTotal Effect

(b) Scenario II: Incremental Change in Daily Exposure

0.00

0.20

0.40

0.60

0.80

1.00

0 5 10 15 20 25 30

Time (day)

Sen

sitiv

ity In

dex

Main EffectTotal Effect

(c) Scenario III: Cumulative Exposure

0.00

0.20

0.40

0.60

0.80

1.00

0 5 10 15 20 25 30Time (day)

Sen

sitiv

ity In

dex

Main EffectTotal Effect

Figure 5-21. Average Main and Total Effects and Corresponding Uncertainty Ranges Based on

500 Bootstrap Simulations for Residue Decay Rate for: (a) Scenario I-Daily Exposure; (b) Scenario II-Incremental Change in Daily Exposure; and (c) Scenario III-Cumulative Exposure.

.

93

accounting for both the linear effect and at least some portion of the nonlinear effect. However,

the correlation methods cannot explicitly account for interactions, and the functional form of the

regression models used here did not include interaction terms. Thus, the correlation and

regression case studies illustrate a distinction perhaps between nonlinear and linear effects

attributable to individual inputs, but do not account for more general forms of interaction as does

Sobol’s method.

Figure 5-21 shows results for the fraction of chemical available for transfer. In general,

these results are qualitatively similar to those from other methods. For example, the sensitivity

of daily exposure to FTR decreases with time, the sensitivity of incremental changes in daily

exposure remains approximately constant at a relatively high sensitivity index, and the sensitivity

of cumulative exposure is relatively low and decreases with time. For the latter, the results by

the end of the month are approximately insignificant, as is the case based on other methods. The

large difference in the total and main effect for Scenario II implies a significant interaction or

nonlinear effect. For the correlation and regression methods, a significant difference was

observed between the rank and sample based methods, which also implies that there is some type

of nonlinear effect. Thus, the comparison of results among these methods is qualitatively

consistent.

5.5.3 Ranking for Selected Inputs

At each time step, inputs were ranked based on the relative magnitude of mean total

effects. Each mean value represents the arithmetic average of 500 bootstrap estimates for the

sensitivity index. The rank of one is assigned to an input with highest mean total effect. The total

effects of inputs accounts for all possible interactions, but additional information would be

needed in order to more specifically determine which specific interactions are contribution the

94

(a) Residue Decay Rate

0

5

10

15

20

25

1 5 10 15 20 25 30

Time (day)

Ran

kScenario I Scenario II Scenario III

(b) Fraction of Chemicals Available for Transfer

0

5

10

15

20

25

1 5 10 15 20 25 30Time (day)

Ran

k

Scenario I Scenario II Scenario III

Figure 5-22. Variation of Ranks with Respect to Time for the Three Temporal Scenarios for: (a) Residue Decay Rate; and (b) Fraction of Chemicals Available for Transfer Based on the Sobol’s

Method.

most to the total interaction effect. The total effects are more reliable than the main effects in

order to investigate the overall effect of each single input on the output. Typically, a similar

pattern with respect to change in ranks of inputs compared to correlation and regression-based

methods was observed.

Figure 5-22 (a) shows variation in ranks of DR with respect to time for the three temporal

scenarios. DR typically showed lower sensitivity earlier in the month, while its sensitivity

increased later in the month for the three temporal scenarios. For example, while DR had a range

of ranks between 2 and 16 in the first 15 days of the model simulation, it was typically identified

among the top two sensitive inputs through out the rest of the month in Scenario II. Thus, the

Lower Rank Implies Greater Sensitivity

95

key similarities in these results versus those from correlation, regression, and ANOVA methods

are the typical increase in importance for Scenarios I and III as time elapses. However, for

Scenario II, there is some qualitative difference in results. For the other three sets of methods,

there was typically a peak in the sensitivity index at approximately the 5th to 10th day. However,

this peak is much less evident based on the results from Sobol’s method. The overall insight that

DR is important for Scenario II only late in the month is consistent based on results from all of

the methods.

For FTR, the results from Sobol’s method imply that early in the month this input is

important for all scenarios, that this input tends to remain important for Scenario II throughout

the month, and that this input tends to decrease in importance for Scenarios I and II at about the

6th or 7th day. This result is most similar to that obtained based on regression analysis, although

in regression analysis the importance of this input with respect to Scenario I appears to be higher.

The correlation and ANOVA results also imply that early in the month this input decreases in

importance for Scenario I, but also imply that this input is unimportant for Scenario III by

approximately the 10th day. However, it is possible that an input ranked as high as 5th, as implied

by the Sobol’s results for Scenario III late in the month, might contribute very little to the overall

variance in the model output. Thus, a rank of five might not imply any significant magnitude of

importance.

5.5.4 Comparison of Mean Ranks

This section presents the mean rank associated with each input within the time period of

the model simulation. The mean rank represents the arithmetic average of ranks for each input.

The range of ranks also is given for each input. The range of ranks represents minimum and

maximum ranks for the input within 30 days of model simulation. Mean ranks for inputs are

96

estimated based on the relative magnitude of main and total effects in 30 days of model

simulation. Mean ranks based on the two measures are compared for the three exposure

scenarios.

Figure 5-23 shows the mean rank and corresponding range of ranks based on main and

total effects for each input. The input abbreviations are given in Table A.1 in Appendix A. There

were two key differences between the mean rankings based on the main and total effects of

inputs: (1) mean ranks based on the two sensitivity indices were different for the top sensitive

inputs; and (2) range of ranks associated with each input based on the main effects was wider

compared to that based on the total effects.

In general, the rankings obtained based on the total effects have significant concordance

with the results obtained from other methods. For example, inputs such as PWB, AM, Area, DR,

and FTR are among those with average ranks in the top five for Scenarios I and III, and this is

typically true based on results from other methods. Similarly, inputs that are identified to have

low average ranks (large values of rank), such as Cair, Cb, RN/T, Height, PAI, Dinh, and others, are

consistently identified as unimportant by all of the methods.

However, there are some potentially significant differences in rankings based upon

Sobol’s method depending on whether only main effects are considered or whether interaction

effects are considered. For example, probability of washing body (PB) had a mean rank of 12

with a range of ranks between 2 and 23 in Scenario II based on the relative magnitude of the

main effects. However, PB had a mean rank of 3 with a range of ranks between 1 and 9 based on

the relative magnitude of the total effects. Because in Scenario II the contribution of inputs to the

97

(a) Ranks Based on Main Effects

0

5

10

15

20

25

Cai

r

Cb

WB k

PH

PW

B

AM

Are

a

RN

/T

Hei

ght

DR

PA

I

Din

h

HW

BW LH LB E

f

TCha

nd

TCbo

dy

ED

targ

et

ED

non-

targ

et

FTR

FBU HF

Inputs

Mea

n R

anks

Scenario I Scenario II Scenario III

(b) Ranks Based on Total Effects

0

5

10

15

20

25

Cai

r

Cb

WB k

PH

PW

B

AM

Are

a

RN

/T

Hei

ght

DR

PA

I

Din

h

HW

BW LH LB E

f

TCha

nd

TCbo

dy

ED

targ

et

ED

non-

targ

et

FTR

FBU HF

Inputs

Mea

n R

anks

Scenario I Scenario II Scenario III

Figure 5-23. Comparison of Mean Ranks and Range of Ranks Based on the Sobol’s Method for Inputs in 30 Days for the Three Temporal Scenarios Based on: (a) Main Effects; and (b) Total Effects.

98

output variance via their main effects was relatively low, and hence, inputs mostly contributed to

the output variance via their nonlinear or interaction effects, a ranking based on the relative

magnitude of the main effects provides misleading results. The differences in rankings of inputs

in the other two exposure scenarios were not that sensitive to the choice of the sensitivity index

as a measure of ranking because the contribution of the main effects to the output variance were

substantially larger in those two scenarios.

Based on the mean total effects, on average, the most sensitive input in Scenario I was

probability of washing body (PWB) with a mean rank of 1.7. The range of ranks for this input was

between 1 and 2. Inputs with secondary sensitivity with respect to daily exposures were residue

decay rate (DR), fraction of chemicals for transfer (FTR), amount of chemical applied (AM), and

body weight (WB). Average ranks for these inputs within 30 days of model simulation were

between 2.6 and 4.6.

In Scenario II, FTR was identified as the most sensitive input with a mean rank of 1.3.

Transfer coefficient via body (TCBody), probability of washing body (PWB), residue decay rate

(DR), exposure duration at target area (EDTarget), and body weight (WB) were identified as inputs

with secondary sensitivity with mean ranks between 3.0 and 4.6.

In Scenario III, PWB was identified as the most sensitive input throughout the entire

month. Inputs with secondary sensitivity included DR, AM, Area, and WB with mean ranks

between 3.0 and 6.0.

5.5.5 Summary of Results Based on Sobol’s Method

Overall, results based on Sobol’s method are qualitatively consistent with those of other

methods. However, Sobol’s method offers an advantage of providing insight into the difference

between the main, linear effect of an input and the total effect that includes all of the possible

99

interactions of a given input with all other inputs. A comparison of the total and main effects

provides some insight into the behavior of the model and the significance of each input when

interactions are accounted for. The Sobol’s algorithm includes a method for reporting results as

a contribution to variance which is a convenient way to summarize and communicate results.

Sobol’s method tends to be much more computationally intensive than methods evaluated in

previous sections.

5.6 Fourier Amplitude Sensitivity Test

This section presents the results based upon the use of FAST to characterize the

sensitivity of the selected model outputs for temporal Scenarios I, II, and III to each of the inputs

of the model. Selected inputs the same as those presented in previous sections are considered as

the focus of discussion here and specific details are given.

Comments are made regarding the contribution of inputs to the output variance (Section

5.6.1). Main and total effect sensitivity indices for selected inputs are given and compared

(Section 5.6.2). Changes in ranking of selected inputs with respect to time are presented (Section

5.6.3). Mean ranks and associated range of ranks in 30 days of model simulation for all inputs

are also provided and compared (Section 5.6.4). Summary of the key findings based on FAST

are discussed (Section 5.6.5).

5.6.1 General Insight Regarding Contribution of Inputs to the Output Variance

For each of the three temporal scenarios, FAST was applied to the model at each time

step. Main and Total effects are estimated for all inputs. The main effect of each input represents

the fractional contribution of the input to the output variance and the sum should equal to one for

linear additive models. However, in the case of non-linearity and interactions between inputs

sum of main effects will be less than one. Figure 5-24 shows the sum of main effects for each of

the three temporal scenarios.

100

(a) Scenario I

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30Time (day)

Sum

of M

ain

Effe

cts

(b) Scenario II: Incremental Change In Daily Exposure

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30Time (day)

Sum

of M

ain

Effe

cts

(c) Scenario II: Cumulative Exposure

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30Time (day)

Sum

of M

ain

Effe

cts

Figure 5-24. Variation in Sum of Main Effects of Inputs Based on FAST with Respect to Time for: (a) Daily; (b) Incremental; and (c) Cumulative Total Exposure Scenarios.

Figure 5-24 (a) shows the sum of main effects of inputs at each time step in Scenario I.

On average, approximately 60 percent of the output variance was apportioned to the main effects

of inputs within the first 10 days of the model simulation. However, in the last 10 days of the

101

model simulation, the sum of main effects decreased to values between 35 and 45 percent. Thus,

approximately 40 percent of the output variance was because of the nonlinearity and possible

interactions between inputs within the first 10 days of the model simulation, while this amount

increased to values between 55 and 65 percent within the last 10 days of the model simulation.

This result is approximately similar to that obtained from Sobol’s method as shown in Figure 5-

18(a).

Figure 5-24 (b) shows sum of main effects of inputs at each time step in Scenario II. In

Scenario II, contribution of main effects of inputs to the output variance was substantially lower.

For example, approximately between 10 and 15 percent of the output variance was apportioned

to the main effects of inputs in the last 10 days of the model simulation. This is also consistent

with the results from Sobol’s method.

Figure 5-24 (c) shows sum of main effects of inputs at each time step in Scenario III. . In

Scenario III, the contribution of main effects of inputs to the output variance did not typically

change at different time steps of the model simulation and was approximately 60 percent. These

findings were consistent with the results shown in Figure 5-8 for the variation of R2 values with

respect to time and those based on the Sobol’s method in Figure 5-18(c).

5.6.2 Sensitivity Indices

For each of the three temporal scenarios, main and total effect sensitivity indices based

on FAST were estimated at each day for all inputs. Thus, 30 sets of main and total effects were

estimated at each time step of the model simulation. Typically, sensitivity indices varied in time

in the three temporal scenarios. Two examples are provided.

102

Figure 5-25 shows variation in main and total effects with respect to time for residue

decay rate (DR). These results are qualitatively similar to those from all other methods, including

Sobol’s method.

For all three scenarios, DR typically has smaller main and total effects in the earlier time

steps compared to the later time steps of the model simulation. Moreover, the difference between

main effects and total effects representing contribution of DR to the output variance in form of

interactions with other inputs increased in later time steps.

For Scenario I, the difference between total and main effects of DR varied between 0 and

0.16 in the first 10 days of the model simulation in Scenario I, while this value ranged between

0.35 and 0.45 in the last 10 days of the model simulation. Thus, approximately 16% of the

output variance can be apportioned to interaction between DR and other inputs at Day 10 and 40

percent at Day 30. The main effect of 0.2 at Day 30 is comparable to, although slightly lower,

than the main effect estimated by Sobol’s method.

In Scenario II, DR typically contributed to the output variance via interactions with other

inputs since the corresponding main effects were negligible in each time step. For example, on

Day 30, although 59% of the output variance was apportioned to overall effect of DR, only 3% of

this amount was because of the direct contribution of DR to the output variance via its main

effect. Thus, 97% of the overall contribution of DR to the output variance was because of

interaction of DR with other inputs in the model. However, individual interaction effects are not

quantifiable based on the sensitivity index, and thus it is not possible to have additional detail on

which combinations of inputs are contributing to the large interaction effect.

Figure 5-26 shows variation in main and total effects with respect to time for fraction of

chemicals available for transfer (FTR). These results are qualitatively similar to those of other

103

(a) Scenario I: Daily Exposure

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30Time (day)

Sen

sitiv

ity In

dex

Main EffectTotal Effect

(b) Scenario II: Incremental Change in Daily Exposure

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30Time (day)

Sen

sitiv

ity In

dex

Main EffectTotal Effect

(c) Scenario III: Cumulative Exposure

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30Time (day)

Sen

sitiv

ity In

dex

Main EffectTotal Effect

Figure 5-25. Main and Total Effects for Residue Decay Rate Based on FAST for: (a) Scenario I-Daily Exposure; (b) Scenario II-Incremental Change in Daily Exposure; and (c) Scenario III-

Cumulative Exposure.

104

(a) Scenario I: Daily Exposure

0

0.2

0.4

0.6

0 5 10 15 20 25 30Time (day)

Sen

sitiv

ity In

dex

Main EffectTotal Effect

(b) Scenario II: Incremental Change in Daily Exposure

0

0.2

0.4

0.6

0 5 10 15 20 25 30Time (day)

Sen

sitiv

ity In

dex

Main EffectTotal Effect

(c) Scenario III: Cumulative Exposure

0

0.2

0.4

0.6

0 5 10 15 20 25 30Time (day)

Sen

sitiv

ity In

dex

Main EffectTotal Effect

Figure 5-26. Main and Total Effects for Fraction of Chemicals Available for Transfer Based on

FAST for: (a) Scenario I-Daily Exposure; (b) Scenario II-Incremental Change in Daily Exposure; and (c) Scenario III-Cumulative Exposure.

105

methods, including Sobol’s method. Typically, the main and total effects associated with FTR

decreased monotonically with respect to time in Scenarios I and III. In Scenario II, although

main effect of FTR decreased in time, the contribution of interactions involving FTR to the output

variance typically increased, and hence, the total effect of the input remained constant around an

approximate value of 0.35.

5.6.3 Ranking for Selected Inputs

At each time step, inputs were ranked based on the relative magnitude of the total effects.

The rank of one is assigned to an input with highest total effect. As discussed for the Sobol’s

method, the total effects of inputs provides a complete characterization of the sensitivity but does

not enable attribution of the specific interactions that are contributing the most to the overall

interaction effect. However, the total effects are more reliable than the main effects in order to

investigate the overall effect of each single input on the output. In Section 5.6.4, mean ranks

based on average main effects and average total effects are compared to evaluate possible

misleading insight with respect to sensitivity based on main effects. Typically, similar pattern

with respect to change in ranks of inputs compared to previous methods was observed.

Figure 5-27 (a) shows variation in ranks of DR with respect to time for the three temporal

scenarios. DR typically showed lower sensitivity earlier in the month, while its sensitivity

increased later in the month for the three temporal scenarios. For example, while DR had a range

of ranks between 16 and 2 in the first 7 days of the model simulation, it was typically identified

among the top two sensitive inputs through out the rest of the month in Scenario II.

Figure 5-27 (b) shows an example in which selected input had relatively higher

sensitivity in the earlier time steps of the model simulation. FTR had higher sensitivity within the

first 5 days of the model simulation compared to the rest of the month for daily and cumulative

106

(a) Residue Decay Rate

0

5

10

15

20

25

1 5 10 15 20 25 30

Time (day)

Ran

kScenario I Scenario II Scenario III

(b) Fraction of Chemicals Available for Transfer

0

5

10

15

20

25

1 5 10 15 20 25 30

Time (day)

Ran

k

Scenario I Scenario II Scenario III

Figure 5-27. Variation of Ranks with Respect to Time for the Three Temporal Scenarios for: (a) Residue Decay Rate; and (b) Fraction of Chemicals Available for Transfer Based on FAST.

exposure scenarios. However, FTR was identified among the top two sensitive inputs through the

whole month in Scenario II.

In general, the results from FAST for the ranking over time of these two selected inputs

were similar to those obtained from other methods. The key insights, such as that DR is

unimportant early in the month but most important late in the month, and that FTR is most

important early in the month and of moderate importance late in the month, are the same based

on all methods.

5.6.4 Comparison of Mean Ranks

This section presents the mean rank associated with each input within the time period of

the model simulation. The mean rank represents the arithmetic average of ranks for each input.

Lower Rank Implies Greater Sensitivity

107

The range of ranks also is given for each input. The range of ranks represents minimum and

maximum ranks for the input within 30 days of model simulation. Mean ranks for inputs are

estimated based on the relative magnitude of main and total effects in 30 days of model

simulation. Mean ranks based on the two measures are compared for the three exposure

scenarios.

Figure 5-28 shows the mean rank and corresponding range of ranks based on main and

total effects for each input. The input abbreviations are given in Table A.1 in Appendix A. There

were two key differences between the mean rankings based on the main and total effects of

inputs: (1) mean ranks based on the two sensitivity indices were different for the top sensitive

inputs; and (2) range of ranks associated with each input based on the main effects was wider

compared to that based on the total effects.

In general, the rankings based on the total effect are similar to those obtained from other

methods. The rankings based on the total effect are more reliable than those based only on the

main effect, because of the effect of nonlinearities and interactions. For example, in Scenario II

probability of washing body (PWB) had a mean rank of 11.5 with a range of ranks between 2 and

22 based on the relative magnitude of the main effects. However, based on the total effects PWB

had a higher mean rank of 4.2 with a narrower range of ranks between 2 and 8. Because in

Scenario II the contribution of inputs to the output variance via their main effects was relatively

low, and hence, inputs mostly contributed to the output variance via their nonlinear or interaction

effects, ranking based on the relative magnitude of the main effects provided misleading results.

The differences in rankings of inputs in the other two exposure scenarios were not as sensitive to

the choice of the sensitivity index as a measure of ranking because the contribution of the main

effects to the output variance were substantially larger in those two scenarios.

108

(a) Rank Based on Main Effects

0.0

5.0

10.0

15.0

20.0

25.0

Cai

r

Cb

WB k

PH

PW

B

AM

Are

a

RN

/T

Hei

ght

DR

PA

I

Din

h

HW

BW LH LB E

f

TCha

nd

TCbo

dy

ED

targ

et

ED

non-

targ

et

FTR

FBU HF

Inputs

Mea

n R

ank

Scenario I Scenario II Scenario III

(b) Ranks Based on Total Effects

0.0

5.0

10.0

15.0

20.0

25.0

Cai

r

Cb

WB k

PH

PW

B

AM

Are

a

RN

/T

Hei

ght

DR

PA

I

Din

h

HW

BW LH LB E

f

TCha

nd

TCbo

dy

ED

targ

et

ED

non-

targ

et

FTR

FBU HF

Inputs

Mea

n R

ank

Scenario I Scenario II Scenario III

Figure 5-28. Comparison of Mean Ranks and Range of Ranks Based on the FAST for Inputs in 30 Days for the Three Temporal Scenarios Based on: (a) Main Effects; and (b) Total Effects.

109

Based on the mean total effects, on average, the most sensitive input in Scenario I was

probability of washing body (PWB) with a mean rank of 2.1. The range of ranks for this input was

between 1 and 3. Inputs with secondary sensitivity with respect to daily exposures were fraction

of chemicals for transfer (FTR), residue decay rate (DR), amount of chemical applied (AM), and

body weight (WB). Average ranks for these inputs within 30 days of model simulation were

between 2.8 and 5.8.

In Scenario II, FTR was identified as the most sensitive input with a mean rank of 1.4.

Residue decay rate (DR), transfer coefficient via body (TCBody), probability of washing body

(PWB), exposure duration at target area (EDTarget), and body weight (WB) were identified as inputs

with secondary sensitivity with mean ranks between 3.5 and 4.8.

In Scenario III, PWB was identified as the most sensitive input with a mean rank of 1.4.

Inputs with secondary sensitivity included DR, AM, Area, and WB with mean ranks between 2.8

and 5.8.

5.6.5 Summary of Results Based on FAST

Summary of key insights and findings based on FAST include:

• The magnitude of main and total sensitivity indices typically changed for inputs with

respect to time.

• Main and Total Effects of inputs with a monthly sampling strategy typically: (1)

declined monotonically; or (2) increased monotonically with respect to time.

• Main effect for inputs with a daily sampling strategy typically declined monotonically

with respect to time.

• Some inputs with a daily sampling strategy had increase in their contribution to the

output variance via later in the month in the form of interactions with other inputs.

110

• Total effects provided a better sensitivity index for ranking inputs specially when

there is substantial non-linearity in the model.

• Results for Scenario II for the top sensitive inputs were typically different from those

in the other two temporal scenarios.

5.7 Comparison of Results for Selected Sensitivity Analysis Methods

This section presents a comparison between the results from selected sensitivity analysis

methods. Several of the selected sampling-based methods are based on either a linearity

assumption (e.g., sample correlation, sample regression) or monotonic relationship between

output and inputs (e.g., rank correlation, rank regression). For the case of ANOVA, which does

not impose any specific model structure assumptions, only the main effects of inputs were

considered in the analysis. Hence, possible interaction effects between inputs were put aside.

This is not an inherent limitation of ANOVA, but is simple a limitation of the case studies. The

use of ANOVA in order to quantify the main effects only was intended to represent a relative

simple application of ANOVA. Both Sobol’s method and FAST account for main effects and

interaction effects. Although there is a limitation in the algorithm used here for FAST, as

described in Chapter 3, the performance of FAST appears to be reasonable in comparison to

other methods, and thus the performance of FAST does not appear to have been significantly

comprised by the presence of frequencies that were not commensurate. Of course, whether

FAST would continue to perform robustly in other model applications with this limitation would

require additional study. This is addressed in the recommendations of Chapter 6.

One key question is with respect to reliability of rankings based on the selected sampling-

based techniques when compared to those based on the variance-based methods such as FAST

and the Sobol’s method. The selected variance-based techniques are global sensitivity analysis

111

methods which do not assume any functional form for the relationship between the output and

inputs and can be applied to models that are non-linear and/or non-monotonic. Among the three

alternative exposure scenarios defined in this chapter, Scenario II provided the most non-linear

response in the model. Typically, R2 values from application of standardized linear regression

based on the sampled and rank transformed data to Scenario II were substantially low in different

time steps of the model simulation. Thus, Scenario II provides an appropriate testbed for

evaluation of selected sensitivity analysis methods.

Figure 5-29 shows mean ranks of inputs in 30 days based on different sensitivity analysis

methods for Scenario II. Range of ranks for each input is also provided. Each range of rank

represents minimum and maximum rank of the input in 30 days of model simulation. Mean ranks

are shown for two categories of inputs: (a) inputs with a monthly sampling frequency; and (b)

inputs with a daily sampling frequency. There are two key differences between the results based

on the sampling-based and variance-based methods: (1) typically, variance-based method

provided lower mean ranks for the most sensitive inputs, particularly those with a monthly

sampling frequency; and (2) the range of ranks for each input based on the variance-based

methods was narrower compared to that based on the sampling-based methods. One example is

provided for each of the inputs with monthly and daily sampling frequencies.

Probability of washing body (PB) is an input with a monthly sampling frequency. The

mean ranks for PB based on the sampling-based methods were between 10.4 and 14.5 with a

112

(a) Inputs with a Monthly Sampling Strategy

0.00

5.00

10.00

15.00

20.00

25.00

Cai

r

Cb

WB k

PH

PW

B

AM

Are

a

RN

/T

Hei

ght

DR

Inputs

Mea

n R

ank

PCA SCA SRA RRA ANOVA Sobol FAST

(b) Inputs with a Daily Sampling Frequency

0.00

5.00

10.00

15.00

20.00

25.00

PA

I

Din

h

HW

BW LH LB E

f

TCha

nd

TCbo

dy

ED

targ

et

ED

non-

targ

et

FTR

FBU HF

Inputs

Mea

n R

ank

PCA SCA SRA RRA ANOVA Sobol FAST

Figure 5-29. Comparison of the Mean Ranks and Corresponding Range of Ranks in 30 Days of Model Simulation Based on Pearson Correlation Analysis (PCA), Spearman Correlation Analysis (SCA), Sample Regression Analysis (SRA), Rank Regression Analysis (RRA), Analysis of Variance (ANOVA), Sobol’s Method, and FAST for: (a) Inputs with Monthly Sampling Frequencies; and (b)

Inputs with Daily Sampling Frequencies in Scenario II.

113

range of ranks between 5 and 25. However, the mean ranks for PB based on FAST and Sobol’s

methods were 3 and 4.2, respectively, with a range of ranks between 1 and 9.

Maximum dermal load via body (LB) is an input with a daily sampling frequency. The

mean ranks for LB based on the sampling-based methods were between 10.0 and 14.9 with

arange of ranks between 2 and 25. However, the mean ranks for PB based on FAST and Sobol’s

methods were 4.6 and 3.5, respectively, with a range of ranks between 2 and 12.

Thus, in the selected examples, there are some differences in results when comparing

methods. However, when looking at the overall set of results, there are more similarities than

differences. For example, the ranks obtained for inputs such as Cair, Cb, k, Ph, AM, Area, RN/T,

Height, PAI, Dinh, LH, Ef, TCBody, EDtarget, EDnon-target, FTR, and HF were approximately the same

among all methods.

When there were differences between methods, the differences were primarily between

Sobol’s method versus correlation, regression, and ANOVA, or between FAST and correlation,

regression, and ANOVA. In these cases, Sobol’s method and FAST produced similar results

even though the results of these two methods differed from the other five methods. Examples

include WB, PWB, DR, and LB. These examples tend to be for inputs that are highly sensitive

according to Sobol’s and FAST methods and less sensitive according to the other methods. This

comparison is most likely because Sobol’s method and FAST do a better job of more completely

characterizing the full relationship between the output and a given input, including complex

interaction effects. The interaction affects are only incompletely characterize by the rank

correlation and regression effects. Although interactions were not considered with ANOVA,

they could be in future work.

114

An overall implication of the comparison for Scenario II, which is perhaps the most

challenging of the three temporal scenarios with regard to sensitivity, is that the two variance-

based methods appear to produce better and more reliable results than the other methods. For

example, using the more conventional methods, one might fail to properly recognize the

importance of some of the inputs, with PWB serving as perhaps the best example.

Based upon the case studies reported here, key findings regarding each of the sensitivity analysis

methods and recommendations for implementation and additional research are provided in the

next chapter.

115

6. CONCLUSIONS AND RECOMMENDATIONS

In this chapter, conclusions drawn from the associated research work and the results are

presented. Key questions that are raised in Chapter 1 are addressed. Suggestions for future

research directions are provided.

6.1 Main Characteristics of the SHEDS Models Relevant to the Process of Choosing Appropriate Sensitivity Analysis Methods

The main characteristics of the SHEDS-Pesticides model included: (1) non-linearity and

interaction between inputs; (2) saturation points; (3) different input types (e.g., continuous versus

categorical); and (4) aggregation and carry-over effects. A simplified version of the SHEDS-

Pesticides model was developed for evaluation of sensitivity analysis methods. The simplified

model retains three of these main characteristics and does not include categorical inputs. An

ideal sensitivity analysis method should be model independent. Specifically, a sensitivity

analysis method should not require any assumptions regarding the functional form of the risk

model and should be applicable to different model formulations. Some methods are considered to

be global and model-independent. However, some methods, including many that are commonly

used, are based upon assumptions regarding the functional form of the model. If a sensitivity

analysis method based upon an assumed functional form of a model is applied to a model with

different characteristics, then the results of the sensitivity analysis may not be valid.

6.2 Available Sensitivity Analysis Methods

Sensitivity analysis methods are typically classified into categories based on their scope,

applicability, and characteristics. We classified sensitivity analysis methods into two categories

of local and global techniques based on Saltelli et al. (2000). Examples of commonly used

sensitivity analysis methods are regression-based techniques (e.g., Helton, 1993), variance

decomposition methods (e.g., Saltelli et al., 1999), and scatter plots (e.g., Kleijnen and Helton,

116

1999; Saltelli et al., 2000; Frey and Patil, 2002). We evaluated a series of seven global sensitivity

analysis methods including Pearson and Spearman correlation coefficients, standardized sample

and rank linear regression, analysis of variance, FAST, and the Sobol’s method via cased studies

with the simplified SHEDS-Pesticides model. Each of these methods has key assumptions that

may or may not conform to the key characteristics of the simplified model. Thus, the case

scenarios in this report provided a testbed for evaluation of selected sensitivity analysis methods

with respect to situations in which the underlying assumptions of each technique may be

challenged by the model.

6.3 Sensitivity Analysis Methods for Application to the SHEDS Models

Table 6-1 summarizes the comparison of evaluated sensitivity analysis methods. Based

on the key criteria, methods should: be able to consider simultaneous variation in inputs; be

computationally efficient; have quantitative measures for ranking key inputs; be reproducible; be

able to apportion the output variance to different model inputs; be model independent; and be

robust in practice. A method is considered as computationally efficient if: (1) statistical software

packages are available for implementation of the method; and (2) the number of model inputs

does not affect the required computational resource. Reproducibility means to what extent

repetition of the calculation procedure leads to the same results. Model independence means the

extent in which the level of additivity of non-linearity of the model influences the correctness of

the results. Robustness means reliability of results when key assumptions of the method are not

met.

117

Table 6-1. Comparison of Various Sensitivity Analysis Methods Sensitivity Analysis Method (a)

Criteria for Comparison (b) PCA SCA SRA RRA ANOVA FAST SobolSimultaneous Variation in Inputs ++ ++ ++ ++ ++ ++ ++ Computational Efficiency ++ ++ ++ ++ +/- +/- +/- Quantitative Measure for Ranking ++ ++ ++ ++ ++ ++ ++ Reproducibility +/- +/- +/- +/- +/- ++ ++ Ability to Apportion the Output Variance

No No +/- +/- +/- ++ ++

Model Independence No No No No Yes Yes Yes Robustness +/- +/- +/- +/- +/- ++ ++ Discrete Inputs - - +/- +/- ++ - Number of Inputs ++ ++ ++ ++ +/- - +/-

a PCA: Pearson Correlation Analysis; SCA: Spearman Correlation Analysis; SRA: Standardized Sample Linear Regression Analysis; RRA: Rank Regression Analysis; ANOVA: Analysis of Variance; FAST: Fourier Amplitude Sensitivity Test

b ++: Completely satisfies the criterion; +/-: Does not completely satisfy the criterion; -: Poorly addresses the criterion

All methods evaluated in this work were able to consider simultaneous variation in

inputs. Sampling-based methods, including correlation and regression based methods, ANOVA

and Sobol’s method are based on Monte Carlo simulation of the model, and hence, incorporate

simultaneous variation of inputs as a part of their sampling strategy. FAST is not based on Monte

Carlo. However, values are generated from the domain of variation of inputs using

transformation functions.

Typically sampling-based methods are computationally efficient. Most of these methods

are built-in features of commonly used software tools such as @RiskTM and Crystal BallTM and

also are available in statistical software packages such as SAS® and free software such as R. The

computing resources required for each of these methods are usually proportional to the resources

necessary to perform a typical Monte Carlo probabilistic simulation with the model. With

available computer hardware, the number of model inputs does not substantially affect the

computational resource required for the sampling-based methods. However, for ANOVA the

118

need for defining levels for continuous inputs and classifying sampled values from probability

distributions of inputs into the levels can impose some computational burden for models with

many inputs. There are few software packages for implementation of FAST and the Sobol’s

methods. Although these are powerful methods that offer advantages over commonly used

techniques, their widespread practical application is limited until software becomes available by

which these methods can be easily incorporated into a risk model (Saltelli et al., 2004).

For the case scenarios in this research, algorithms were developed for FAST and Sobol’s

method and coded in Matlab. The algorithm for FAST requires specification of frequencies that

dictate how each input is sampled. As an ideal and desirable specification of frequencies

considered in FAST is that they should be incommensurate. The incommensurate frequencies

typically have an important property that no single frequency in a frequency set can be expressed

as a linear combination of other frequencies (Lu and Mohanty, 2001). If this criterion is not

satisfied by the set of frequencies considered for the inputs, an interference problem will be

introduced. Interference is a problem when information provided by the frequencies is mixed;

hence, the problem could lead to overestimating the sensitivity indices, and hence, the

contribution of inputs to the output variance (Saltelli et al, 2000). In practice, a published table is

used to produce incommensurate frequencies for up to 50 inputs (Cukier et al, 1978; Saltelli et al,

2000). However, there is not a published record of an algorithm for generating a larger number

of incommensurate frequencies for more than 50 inputs.

Because the simplified SHEDS-Pesticides model has 30 time periods for each of the daily

inputs, and because the goal was to generate independent samples in each of the time periods,

there is a need for different frequencies for a given input variable for each time period (day) of

the simulation. Thus, although there are only 14 daily inputs, there is a need to generate 30 × 14

119

(i.e., 420) independent sets of samples. In addition, there are 11 monthly inputs. Thus, the total

number of independently sampled distributions is 431. Thus, there was a need to generate

additional frequencies beyond the set of 50 incommensurate frequencies available in literature.

In order to generate additional frequencies, an algorithm was needed. Since no algorithm was

located in the literature, a simplified algorithm was developed that produces frequencies that are

not repetitive. However, this does not guarantee that the frequencies are incommensurate.

Although the case study results from FAST were comparable to those from Sobol’s method,

even though frequencies that were incommensurate were included as a basis for the simulation,

this does not guarantee that FAST will be robust when applied more generally. For example, it

could simply be that in cases where two or more frequencies were not incommensurate, they

happened to be assigned to inputs that did not have any significant influence on the results.

The implications of the need for incommensurate frequencies for FAST, and the lack of

availability of either data for such frequencies or an algorithm for generating such frequencies

when more than 50 inputs are needed is that more research is needed in order to further develop

or evaluate FAST when many inputs are to be analyzed. For the typical SHEDS model, there

can be many inputs and many different time periods, with possibly different times scales for

various inputs. Thus, a sensitivity analysis method must be able to accommodate a large number

of inputs in order to be applicable to the SHEDS models. This research has represented the first

known application of FAST to a model of this type. Thus, as part of this research, a limitation of

FAST has been identified that does not typically apply to the more routine applications of FAST

reported in the literature. There are two general ways to deal with the limitation on the number

of incommensurate frequencies that are available for FAST. One is to conduct additional

research to develop an algorithm for generating additional such frequencies in whatever numbers

120

are needed for application to a given SHEDS model. A second is to evaluate the robustness of

FAST even if the requirement for incommensurate frequencies is not satisfied, in order to

identify conditions under which the results might be correct as long as interference problems are

minimized.

For Sobol’s method, and in some preliminary analysis with a simple linear model,

Sobol’s method provided accurate sensitivity coefficients. However, in one case in which the

scale of the inputs was changed, the convergence toward accurate values of sensitivity indices

required a substantially larger number of iterations. This comparison suggests that the number of

iterations required for Sobol’s method may be sensitive to the scale of the model inputs or to

whether the variation in the model output is large. In practice, and as shown in Chapter 5,

Sobol’s method provided reasonable results for the simplified SHEDS-Pesticides model in

comparison to other sensitivity analysis methods, and thus a convergence problem was not

suspected there. The output for the simplified SHEDS-Pesticides model has a large range of

variation relative to its mean. A hypothesis is that in this type of situation, Sobol’s method is able

to converge on accurate sensitivity indices with a moderate number of iterations. However, in

order to address this problem, the output of Sobol’s method was normalized, and the normalized

estimates of contribution to variance were found to converge to the correct value more quickly

than an estimate based on the raw sensitivity indices. Thus, the use of normalized sensitivity

measures appears to facilitate more rapid convergence, thereby reducing the needed sample size.

However, the properties of the normalization with respect to the necessary sample size require

additional assessment in order to prepare guidance on how to choose a sample size for Sobol’s

method.

121

There may be variation in the rate at which Sobol's method converges on accurate

values of the sensitivity coefficients, possibly depending on the scale of the inputs or depending

on whether the range of variability in the output is small relative to its mean. However, in future

work the hypothesis should be tested as to whether Sobol's method may require more iterations

to converge on sensitivity coefficients depending on the scale of model inputs, on the variance of

the model output relative to its mean, or the normalization method used to estimate the

contribution to variance.

Sobol’s method is not computationally efficient for models with many inputs. Full sets of

first plus total effect sensitivity indices can be estimated at the cost of )22( +×× knt

simulations, where t is the number of time steps considered in the analysis (e.g., 30 for the case

scenarios in this research), n is the number of bootstrap simulations for estimating the confidence

intervals for sensitivity indices, and k is the number of model inputs. If sensitivity indices for

other effects such as two-way and three-way interactions between inputs are to be estimated, the

computational cost will increase drastically.

All of the selected methods for sensitivity analysis provide quantitative measures for

ranking inputs. The ability to produce quantitative rankings and the ability to evaluate the

statistical significance of the rankings are useful to identify the relative importance of inputs and

the confidence that should be imputed to the rankings. Some methods produce more useful

measures by which to discriminate the importance among similarly ranked inputs. For example,

correlation and regression-based methods provide confidence intervals for sensitivity indices

based on which statistically significant difference between indices associated with different

inputs can be inferred. Sobol’s method includes bootstrap simulation, and hence, is able to

122

identify the standard error associated with the sensitivity indices. However, for FAST no

criterion was identified to evaluate whether sensitivity indices were statistically different.

Results based on the sampling-based techniques are sensitive to the simulation sample

size. With small sample sizes, the results may change based on different set of samples. Frey et

al. (2003) show some case studies in which sensitivity analysis results based on ANOVA

changed substantially when alternative set of samples were used for the analysis. Variance-based

techniques evaluated in this research were typically reproducible. FAST produces the same

mapping from the input domain to the model output space when the number of inputs does not

change. The Sobol’s method incorporates the application of the bootstrap technique in estimation

of sensitivity indices, and hence, can quantify possible uncertainty in those estimates due to the

sample size.

The ability to apportion the output variance to individual inputs and their combinatory

effects has been considered as the key objective of this research. Among the selected sensitivity

analysis methods, Sobol’s method and FAST directly provide insight with respect to this

objective. Sensitivity indices based on these two methods represent the contribution of the

selected individual input or combination of inputs to the output variance. For example, a main

effect of 0.5 for the input x1 indicates that 50% of the output variance can be apportioned to x1.

Correlation-based methods are not able to provide insight with regard to the contribution to

variance. However, regression-based methods can provide some insight in this regard. The

coefficient of determination, R2, represents the amount of output variability that can be captured

by a regression or ANOVA approach. Furthermore, some measure of variation in the output,

such as sum of squares of the output values, can be apportioned to individual terms (e.g., input

terms, interaction terms, or any nonlinear term). However, this issue was not investigated here.

123

The evaluated sampling-based methods are not typically model-independent. These

methods are based upon assumptions regarding the functional form of the model. For example,

sample correlation analysis or linear regression is based upon the assumption of a linear model.

Rank correlation coefficients or rank regression are based upon the assumption of a monotonic

model. ANOVA and the two variance-based methods are model-independent. If a sensitivity

analysis method based upon an assumed functional form of a model is applied to a model with

different characteristics, then the results of the sensitivity analysis may not be valid. For

example, if method based upon linearity, such as sample correlation analysis, is applied to a

nonlinear or non-monotonic model, then the insights regarding sensitivity could be inaccurate or

invalid.

Results based on the sampling-based methods of correlation and regressions are typically

perceived, on a theoretical basis, not to be robust because these methods are not model

independent. However, in practice, the results obtained from these methods, and particularly for

correlation coefficients, were comparable to those from ANOVA, which is considered to be a

model-independent technique. Thus, in practice, with at least some kinds of models, correlation

and regression models can enable useful insights regarding sensitivity. For ANOVA, although

this method does not assume any functional form, results can be sensitive to the definition of

levels and the number of data within each level after classification of sampled data (Neter et al.,

1996; Kleijnen and Helton, 1999).

The emphasis of this project has been on evaluation of methods that can provide insight

with respect to contribution of inputs to the output variance. FAST and the Sobol’s methods are

among the variance-based techniques that can apportion the output variance to individual effects

of inputs along with their combinatory contributions. However, and due to computationally

124

intensiveness of these two methods, simpler methods can be used as screening techniques in the

earlier stage of the analysis to identify those inputs that are typically unimportant. Our case study

results typically showed that all methods had agreement in identification of unimportant inputs.

Such inputs with a small percentage of contribution to the output variance can be frozen to any

value within their range of variation (e.g., mean). This process contributes substantially to the

model simplification and applicability of resource intensive methods such as FAST and Sobol’s

method to models with several inputs.

In general, five of the methods, including sample correlation, rank correlation, sample

regression, rank regression, and ANOVA provided results that were comparable to each other.

With regard to correlation and regression, the two rank-based methods are more robust and

reliable than the two sample-based methods. The differences in results between rank and

sample-based methods are somewhat analogous to the comparison of total versus main effect in

methods such as Sobol’s and FAST; however, they are not as general in capturing interaction

effects as are the latter two methods. A comparison of rank versus sample based results typically

provides some insight regarding the importance of nonlinearity in evaluating sensitivity results,

as was illustrated in several case studies.

Both Sobol’s and FAST methods produced similar results to each other, but tended to

differ from the other methods. In particular, for Scenario II, which was perhaps the most

challenging testbed for sensitivity analysis methods because of the large importance of

interaction effects, Sobol’s and FAST methods identified some inputs as more important than did

the other five methods. This difference is most likely because Sobol’s and FAST methods are

more capable of capturing the effect of interactions than are the other methods. However, the

125

application of regression and ANOVA in the case studies here did not take into account

interaction effects, which could be included.

The regression and ANOVA methodologyies have ways of dealing with user-specified

interaction effects. In the case of regression, the interaction effect has to be evaluated with

respect to a particular functional form of a regression model, whereas for ANOVA it is not

necessary to specify a specific functional form. However, even if interaction effects are

accounted for in regression and ANOVA, a practical advantage of Sobol and FAST is that all

interactions are systematically accounted for as an inherent part of the methodology.

Based on the summary results for comparison of the selected sensitivity analysis methods

in Table 6-1, it follows that Sobol’s method produced the best overall performance. FAST is

also a promising method but the current limitations pertaining to development of frequencies that

are used as the basis for sampling require additional research before this method could be

recommended for routine application to the SHEDS models. The other methods, such as

correlation, regression, and ANOVA, are useful as screening methods and are capable of

providing useful results for models that are approximately linear. However, they do not directly

quantify contribution to variance in the manner that Sobol’s method does.

6.4 Recommendations

Key recommendations are as follows:

(1) Sobol’s method is identified as a promising method for apportioning the variance in a

model output to individual inputs and to the interaction effects among inputs.

Therefore, the application of Sobol’s method to a larger, or more complex, model is

recommended as the next step toward more routine applications of this method to

models such as SHEDS. In future work the hypothesis should be tested as to whether

126

Sobol's method may require more iterations to converge on sensitivity coefficients

depending on the scale of model inputs, on the variance of the model output relative

to its mean, or the normalization method used to estimate the contribution to variance.

Furthermore, there is a need for guidance on how to choose a sample size for SHEDS.

There are also some applicability issues that must be addressed, that are covered in

other recommendations below, that are general for many methods including Sobol’s.

(2) The correlation, regression, and ANOVA methods were found to produce comparable

results in most cases, and they enabled insight into complex temporal issues,

nonlinearities, and competition among inputs with respect to importance. These

methods appear to be useful in practice for identifying inputs that do not matter and

thus could be used as a screening step prior to application of a more refined method

such as Sobol’s. Furthermore, a comparison of sample versus rank-based results for

either correlation or regression provides insight into the importance of nonlinearity in

the model and implications for sensitivity. An area of research for regression-based

methods is to explore the use of sum of squares of the output values, in order to

apportioned the response of the output to individual terms (e.g., input terms,

interaction terms, or any nonlinear term) in the context of the SHEDS model.

ANOVA could be extended to include interaction effects in a model independent

manner, implying some advantage over the inherently model-dependent approach of

correlation or regression techniques. In general, it can be useful to apply and

compare results from sample and rank based methods, or to apply and compare

ANOVA results with one of the correlation or regression based approaches. Thus,

127

were possible to apply two or more methods, these methods can be used in

combination or as screening methods.

(3) We did not consider any dependency structure between inputs in the simplified

SHEDS-Pesticides model. However, for cases that inputs have dependency structures

such as correlation, sensitivity analysis methods should be evaluated and, if

necessary, modified in order to incorporate such dependencies in the analysis.

Sensitivity analysis methods should be evaluated with respect to their functionality in

the case of inputs with dependency. Approaches should be explored for quantification

of different dependency structures between inputs and possible modification of

sensitivity analysis methods for such cases.

(4) For the selected variance-based techniques, including Sobol’s method and FAST,

algorithms were provided for estimation of first and total effects of each input.

Although the total effects of inputs are more reliable than the first-order (main

effects) indices, they do not provide a complete characterization of the sensitivity in

that they do not enable a full explanation of the major cause of interactions. Thus,

optimized algorithms should be provided in order to break down the total effect into

interaction effects between inputs, such as pairwise, three-way combinations, and so

on.

(5) The simplified SHEDS-Pesticides model assumes that inputs are the only sources of

variability in the model, and hence, the whole output variance can be apportioned to

inputs. However, there may be stochastic sources of variability in the model that are

not included in the form of an input in the model. For example, data might be

sampled from a database for joint values of several inputs, in order to capture intra-

128

individual dependencies. Selected variance-based sensitivity analysis methods may

fail to provide accurate estimates of sensitivity indices specifically with respect to the

total effect of each input, when stochastic sources of variability are available in the

model. Thus, these methods should be evaluated and modified as needed to be

applicable to models with such sources of variation in the response.

(6) FAST is a promising method for estimating contribution to variance. However,

additional research is recommended in several areas before this method can be

considered for application to models such as SHEDS. One is to conduct additional

research to develop an algorithm for generating additional such frequencies in

whatever numbers are needed for application to a given SHEDS model. A second is

to evaluate the robustness of FAST even if the requirement for incommensurate

frequencies is not satisfied, if a suitable algorithm for generating the needed number

of incommensurate frequencies is not achieved, in order to identify conditions under

which the results might be correct as long as interference problems are minimized. A

third is to evaluate the applicability of FAST with respect to model inputs that are

categorical or discrete.

(7) Future work could explore normalization algorithms for summarizing the fractional

contribution of each input or interaction effect to the sum of sensitivity indices, and

compare the results with those obtained from Sobol’s method or FAST, in order to

determine whether there are practical alternatives to Sobol’s method that might be

more readily available or easier to apply.

(8) This study has focused on the objective of quantifying the contribution to variance of

a model output. If the assessment objective changes, then the choice of an

129

appropriate sensitivity analysis method can also change. For example, if the objective

is to identify key factors influencing high exposures, then a method such as

Categorical and Regression Trees (CART) should be considered. In general, the

results and findings here should not be interpreted to preclude the use of other

methods that might be more appropriate to a particular assessment objective.

131

REFERENCES

Agro, K. E., C. A. Bradley, and N. Mittmann (1997), "Sensitivity Analysis in Health Economic

and Pharmacoeconomic Studies - an Appraisal of the Literature," Pharmacoeconomics,

11(1):75-88.

Archer, G., A. Saltelli, and I. M. Sobol (1997), “Sensitivity Measures, ANOVA-Like

Techniques, and the Use of Bootstrap,” Jurnal of Statistical Computing and Simulation,

58(1): 99-120.

Baker, S., D. Ponniah, and B. Klubek (1999), “Survey of Risk Management in Major UK

Companies,” Journal of Food Protection, 62(9): 1050-1053.

Baniotopoulos, C. C. (1991), "A Contribution to the Sensitivity Analysis of the Sea-Bed

Structure Interaction Problem for Underwater Pipelines," Computers & Structures,

40(6):1421-1427.

Beck, M. B., J. R. Ravetz, and L. A. Mulkey (1997), "On the Problem of Model Validation for

Predictive Exposure Assessments," Stochastic Hydrology and Hydraulics, 11(3):229-254.

Cheng, T. C. E. (1991), "EPQ with Process Capability and Quality Assurance Considerations,"

Journal of the Operational Research Society, 42(8):713-720.

Cukier, R. I., C. M. Fortuin K. E. Shuler, et al. (1973), ”Study of the Sensitivity of Coupled

Reaction Systems to Uncertainties in Rate Coefficients: Part I Theory,” Journal of

Chemical Physics, 59, 3873-3878.

Cukier, R. I., J. H. Schaibly, and K. E. Shuler (1975), “Study of the Sensitivity of Coupled

Reaction Systems to Uncertainties in Rate Coefficients: Part III Analysis of

Approximation,” Journal of Chemical Physics, 63(5), 1140-1149.

132

Cukier R. I. , H. B. Levine, and K. E. Shuler (1978), “Nonlinear Sensitivity Analysis of Multi-

Parameter Model Systems,” Journal of Computational Physics, 26(1): 1-42.

Cullen, A. C, and H. C. Frey (1999). Probabilistic Techniques in Exposure Assessment. Plenum

Press, New York, USA.

DeGroot, M. H. (1986). Probability and Statistics, 2nd Edition. Addison-Wesley, MA.

Devore, J. L., and R. Peck (1996). Statistics: The Exploration and Analysis of Data. 3rd Edition.

Brooks/Cole Publishing Company: London, England.

Devroye, L. (1986). Non-Uniform Random Variant Generation. Springer-Verlag, New York.

Doll, J. D., and D. L. Freeman (1986), “Randomly Exact Methods,” Science, 234:1356-1360.

Fishman, G. S. (1996). Monte Carlo: Concepts, Algorithms, and Applications. Springer Verlag,

New York.

Fraedrich, D., and A. Gildberg (2000), “A Methodological Framework for the Validation of

Predictive Simulations,” European Journal of Operational Research, 124(1): 55-62.

Frey, H. C. (1992), “Quantitative Analysis of Uncertainty and Variability in Environmental

Policy Making,” American Association for Advancement of Science, Washington DC.

Frey, H. C. (1997). Variability and uncertainty in highway vehicle emission factors. In Emission

Inventory: Planning for the Future (held October 28-30 in Research Triangle Park, NC),

Air and Waste Management Association, Pittsburgh, Pennsylvania, pp.208-219.

Frey, H. C., R. Baharkivar, and J. Zheng (1999). Quantitative analysis of variability and

uncertainty in emission estimations. Final Report, Prepared by North Carolina State

University for Office of Air Quality Planning and Standards, U. S. Environmental

Protection Agency, Research Triangle Park, NC.

133

Frey, H. C., A. Mokhtari, and J. Zheng (2004). Recommended practice regarding selection,

application, and interpretation of sensitivity analysis methods applied to food safety

process risk models. Prepared by North Carolina State University for Office of Risk

Assessment and Cost-Benefit Analysis, U.S. Department of Agriculture. Washington DC.

www.ce.ncsu.edu/risk.

Frey, H. C., and R. Patil (2002), “Identification and Review of Sensitivity Analysis Methods,”

Risk Analysis, 22(3): 553-577.

Frey, H. C., D. S. Rhodes (1996), “Characterizing, Simulating and Analyzing Variability and

Uncertainty: An Illustration of Methods Using an Air Toxics Emissions Example,”

Human and Ecological Risk Assessment, 2(4), 762-797.

Frey, H. C., J. Zheng (2000). Methods and example case study for analysis of variability and

uncertainty in emission estimation (AUVEE). Prepared by North Carolina State

University for Office of Air Quality Planning and Standards, U.S. Environmental

Protection Agency, Research Triangle Park, NC.

Frey, H. C., J. Zheng (2002), “Probabilistic Analysis of Driving Cycle-Based Highway Vehicle

Emission Factors,” Environmental Science and Technology, 36(23), 5184-5191.

Gardiner, C. W. (1983). Handbook of Stochastic Methods for Physics, Chemistry and the Natural

Sciences. Springer-Verlag, New York.

Gibbons, J. D. (1985). Nonparametric Statistical Inference. Marcel Dekker, Inc, New York and

Basel, 2nd edition.

Hahn, G. J., and S. S. Shapiro (1967). Statistical Models in Engineering. Wiley Classics Library,

John Wiley and Sons: New York.

134

Hamed, M. M., J. P. Conte, and P. B. Bedient (1995), “Probabilistic Screening Tool for Ground-

Water Contaminant Assessment,” Journal of Environmental Engineering, 121(11):767-

775.

Harrison, K. W. (2002). Environmental and Water Resource Decision Making Under

Uncertainty. Ph.D. Dissertation, North Carolina State University.

Helton, J. C. (1993), "Uncertainty and Sensitivity Analysis Techniques for Use in Performance

Assessment for Radioactive-Waste Disposal," Reliability Engineering and System Safety,

42(2-3): 327-367.

Helton, J. C., and R. J. Breeding (1993), "Evaluation of Reactor Accident Safety Goals,"

Reliability Engineering & System Safety, 39(2):129-158.

Iman, R. L., and W. J. Conover (1980), “Small Sample Sensitivity Analysis Techniques for

Computer Models, with an Application to Risk Assessment,” Communications in

Statistics, Part A. Theory and Methods, 17:1749-1842, 1980.

Isukapalli, S. S., A. Roy, P. G., and P. G. Georgopoulos (1998). Stochastic response surface

methods (SRSMs) for uncertainty propagation: Application to environmental and

biological systems. Online report. Rutgers University, Piscataway, NJ.

http://www.ccl.rutgers.edu/~ssi/srsmreport/index.html

Jones, R. N. (2000),”Analyzing the Risk of Climate Change Using an Irrigation Demand

Model,” Climate Research, 14(2): 89-100.

Karamchandani, A., and C. A. Cornel (1992), “Sensitivity Estimation Within First and Second

Order Reliability Methods,” Structural Safety, 11(2):95-107.

Kalos, M. H., and P. A. Whitlock (1986). Monte Carlo Methods : Basics. John Wiley & Sons,

New York.

135

Kendall, M. G., and J. D. Gibbons (1990). Rank Correlation Methods. Fifth Edition. Charles

Griffin: London.

Kewley, R. H., M. J. Embrechts, and C. Breneman (2000), "Data Strip Mining for the Virtual

Design of Pharmaceuticals with Neural Networks," IEEE Transactions on Neural

Networks, 11(3):668-679.

Kini, M. D., H. C. Frey (1997). Probabilistic evaluation of mobile source air pollution: Volume

1, Probabilistic modeling of exhaust emissions from light duty gasoline vehicles.

Prepared by North Carolina State University for Center for Transportation and the

Environment, Raleigh, December.

Kleijnen, J. P. C. (1995), “Verification and Validation of Simulation-Models,” European Journal

of Operational Research, 82(1): 145-162.

Kleijnen, J. P. C., and J. C. Helton (1999), “Statistical Analyses of Scatterplots to Identify

Important Factors in Large-Scale Simulations, 1: Review and Comparison of

Techniques,” Reliability Engineering and System Safety, 65(2): 147-185.

Kleijnen, J. P. C, and R. G. Sargent (2000), “A Methodology for Fitting and Validating

Metamodels in Simulation,” European Journal of Operational Research, 120(1): 14-29.

Krishnaiah, P. R. (1981). Analysis of Variance. Elsevier: New York.

Limat, S., M. C. Woronoff-Lemsi, E. Deconinck, et al. (2000), "Cost-Effectiveness of Cd34(+)

Dose in Peripheral Blood Progenitor Cell Transplantation for Non-Hodgkin's Lymphoma

Patients: a Single Center Study," Bone Marrow Transplantation, 25(9):997-1002.

Loh, W. H (1996), “On Latin Hypercube Sampling,” Annals of Statistics, 24(5): 2058-2080.

Lu, R., Y. Luo, and J. P. Conte (1994), “Reliability Evaluation of Reinforced Concrete Beam,”

Structural Safety, 14:277-298.

136

Manheim, L. M. (1998), "Health Services Research Clinical Trials: Issues in the Evaluation of

Economic Costs and Benefits," Controlled Clinical Trials, 19(2):149-158.

McCamley, F., and R. K. Rudel (1995), "Graphical Sensitivity Analysis for Generalized

Stochastic Dominance," Journal of Agricultural and Resource Economics, 20(2):403-

403.

Mokhtari, A., and H. C. Frey (2005). “Recommended Practice Regarding Selection of Sensitivity

Analysis Methods Applied to Microbial Food Safety Models,” Human and Ecological

Risk Assessment, 11(3): 1-16.

Morgan, M. G., and M. Henrion (1990). Uncertainty: A Guide to Dealing with Uncertainty in

Quantitative Risk and Policy Analysis. New York: Cambridge University Press.

Morgan, M. G., S. C. Morrise, M. Henrion, et al. (1984), “Technical Uncertainty in Quantitative

Policy Analysis- A Sulfur Air Pollution Example,” Risk Analysis, 4(3): 201-213.

Neter, J., M. H. Kutner, C. J. Nachtsheim, et al. (1996). Applied Linear Statistical Models,

Fourth Edition. McGraw-Hill, Chicago, IL, USA.

Oh, B. H., and I. H. Yang (2000), "Sensitivity Analysis of Time-Dependent Behavior in PSC

Box Girder Bridges," Journal of Structural Engineering-ASCE, 126(2):171-179.

Phillips, A., D. Janies, and W. Wheeler (2000), “Multiple Sequence Alignment in Phylogenetic

Analysis,” Molecular Phylogenetics and Evolution, 16(3):317-330.

Pollak, A. K., P. Bhave, J. Heiken, et al. (1999). Investigation of emission factors in the

California EMFAC7G Model, PB99-14718INZ. Prepared by ENVIRON International

Corp, Novato, CA, for coordinating Research Council, Atlanta, GA.

Price, P. S., F. C. Chaisson, M. Koontz, et al. (2003). Construction of a Comprehensive

Chemical exposure framework using person oriented modeling. Developed for the

137

Exposure Technical Implementation Panel, American Chemistry Council, Contract #

1388. Environmental and Occupational Health Science Institute, Rutgers University.

Rubinstein, R. Y. (1981). Simulation and the Monte Carlo Method. John Wiley & Sons, New

York.

Saltelli, A., and R. Bolado (1998), “An Alternative Way to Compute Amplitude Sensitivity Test

(FAST),” Computational Statistics and Data Analysis, 26: 445-460.

Saltelli, A., S. Tarantola, and K. P. Chan (1999), “A Quantitative Model-Independent Method for

Global Sensitivity Analysis of Model Output,” Technometrics, 41, 39-56.

Saltelli, A., K. Chan, and M. Scott (2000). Sensitivity Analysis: Probability and Statistics Series.

John Wiley & Sons, New York, USA.

Saltelli, A. (2002), “Making Best Use of Model Evaluation to Compute Sensitivity Indices,”

Computer Physics Communications, 145: 280-297.

Schaibly, J. H., and K. E. Shuler (1973), “Study of the Sensitivity of Coupled Reaction Systems

to Uncertainties in Rate Coefficients. Part II, Applications,” Journal of Chemical Physics,

59(6): 3879-3888.

Siegel, S., and N. J. Castellan (1988). Nonparametric Statistics for the Behavioural Sciences (2nd

edn), McGraw-Hill, New York.

Sobol, I.M. (1993), “Sensitivity Estimates for Nonlinear Mathematical Models,” Mathematical

Modeling and Computation, 1(4):407-414.

Sobol, I. M. (1994). A Primer for the Monte Carlo Method. CRC Press.

Stein, M. (1987), ”Large Sample Properties of Simulations Using Latin Hypercube Sampling,”

Technometrics, 29(2):143-151.

138

U.S. EPA (1996),”Summary Report for the Workshop on Monte Carlo Analysis,” EPA/630/R-

96/010, Risk Assessment Forum, Washington, DC.

U.S. EPA (1997),”Guiding Principles for Monte Carlo Analysis,” EPA/630/R-97/001, Risk

Assessment Forum, Washington, DC.

U.S. EPA (2000), “First-Generation Multimedia, Multipathway Exposure Models,” NERL

Research Abstract, National Exposure Research Laboratory, Research Triangle Park, NC,

September 2000.

Ward, M. P., and T.E. Carpenter (1996), “Simulation Modeling of the Effect of Climatic Factors

on Bluetongue Virus Infection in Australian Cattle Herds - 1. Model Formulation,

Verification and Validation,” Preventive Veterinary Medicine, 27(1-2):1-12.

Wilson, R., and E. A. Crouch (1981),”Regulations and Carcinogens,” Risk Analysis, 1(1):47-57.

Zartarian, V. G., H. Ozkaynak, J. M. Burke, et al. (2000), “A Modeling Framework for

Estimating Children’s Residential Exposure and Dose to Chlorpyrifos via Dermal

Residue Contact and Non-Dietary Ingestion,” Environmental Health Perspective, 108(4):

505-514.

139

APPENDIX A

A.1 Simplified SHEDS-Pesticide Model

This section provides the key equations in the simplified SHEDS-Pesticides model along

with the probability distributions of model inputs. Figure A.1 shows a schematic diagram of the

simplified model. The final exposure is estimated based on three pathways as:

iIngestioniDermaliInhalationiTotal EEEE ,,,, ++= (A.1) Where, ETotal, i = Total exposure at time step i, i = 1, 2,…,30 (days) [µg/kg]

EInhalation, i = Exposure via inhalation pathway at time step i [µg/kg]

EDermal, i = Exposure via dermal pathway at time step i [µg/kg]

EIngestion, i = Exposure via non-dietary ingestion pathway at time step i [µg/kg]

i = 1, 2,…, 30; represents the number of days since application of the

pesticide

Equations corresponding to each of the pathways including inhalation, dermal, and ingestion are

presented and briefly explained.

Inhalation

Indoor inhalation exposure at each time step is estimated as:

B

iinhBViiInhalation W

DRPAICE ,

,

×××= (A.2)

Where,

EInhalation, i = Exposure via inhalation pathway at time step i [µg/kg]

iC = modeled or measured airborne concentration of pollutant in air at

the ith day [µg/m3]

140

Figure A.1. Schematic Diagram of the Simplified SHEDS Model.

bairi CikExpCC +×−×= )( (A.3)

Cair = Concentration of applied pesticides [µg/m3]

Cb = Background concentration of pesticides [µg/m3]

k = Decay Rate [1/day]

PAI = Physical activity index [-]

Dinh,i = Inhalation Exposure duration time at the ith day [min]

WB = Body weight [kg]

RBV = Basal ventilation rate [m3/min] as:

1440)11.2095.0(434.1 +×

×= BBM

WR (A.4)

Average exposure over exposure duration:

m

EE

m

iiInhalation

Inhalation

∑== 1

,

(A.5)

Where,

141

m = Total duration of exposure considered in the scenario (e.g., m = 30

days)

Dermal

Exposure from dermal pathway consists of exposure from hand and body (non-hand).

Key characteristics of the dermal pathway include: (1) carryover of exposure from prior event or

prior day; (2) adjustment for maximum dermal loading; (3) adjustment for washing; and (4)

adjustment for hand-to-mouth transfer.

Figure A.1 shows that non-dietary exposure pathway is estimated as a fraction of the

dermal pathway via hands using a coefficient representing the fraction of hands with residue

going into mouth at the ith day (i.e., HFi). Thus, the non-dietary exposure is subtracted from the

dermal exposure via hands. Final dermal exposure via hand at ith day (EdHand, i,final) is estimated

as:

)1(,,,, iwidHandfinalidHand HFEE −×= (A.6) Where,

EdHand, i, final = Final dermal exposure via hand [µg/kg]

EdHand,i,w = Dermal exposure via hands adjusted for washing effect [µg/kg]

HFi = Fraction of hands with residue going into mouth at the ith day [-]

×−=

aidHandi

aidHandwidHand EHW

EE

,,

,,,, )1(

(A.7)

Where, HWi = Hand washing removal efficiency at the ith day [-]

PWH = Probability of washing hands in a day [-]

EdHand,i,a = Adjusted dermal exposure via hands at the ith day [µg/kg]

),( ,,,, cidHandLHaidHand EEMinE

i= (A.8)

Where,

iLHE = Maximum dermal exposure via hand [µg/kg]

P<1-PWH (No Washing)

P>= 1-PWH (Washing)

142

B

handFractotaliLH W

SASALHE

i

××= (A.9)

LHi = Maximum dermal load via hand [µg/cm2]

SAhandFrac = Fraction of body surface area associated with hands [-]

SAtotal = Total body surface area [cm2]

54375.035129.0305 Btotal WHSA ××= (A.10)

H = Height [cm]

EdHand,i,c = Dermal exposure via hand corrected for the carryover effect

[µg/kg]

idHandidHandfcidHand EEEE ,1,,, +×= − (A.11) Where, Ef = Fraction of exposure from the prior day considered in the carryover

effect [-]

EdHand,i-1 = Dermal exposure via hands at the time step i-1 [µg/kg]

EdHand,i = Dermal exposure via hands at the time step i [µg/kg]

Where,

B

iettnonHandettnon

B

iettHandiettidHand W

EDTCSRW

EDTCSRE ,argarg,arg,arg

,−− ××

+××

= (A.12)

Where, SRtarget ,i = Residue at target surfaces at the ith day (measured or modeled)

[µg/cm2]

SRnon-target, i = Residue at non-target surfaces at the ith day [µg/cm2]

TChand = Hand dermal transfer coefficient measured [cm2/hr]

EDtarget,, i = Duration of exposure to target site at the ith day [hr]

EDtarget,, i = Duration of exposure to non-target site at the ith day [hr]

143

For surface residue at the ith day:

iRTRAett DFRSR )1(100arg −×××= (A.13-a)

iRTRTNAettnon DFRRSR )1(100 /arg −××××=− (A.13-b)

Where, RA = Application rate [g/m2]

RA = AreaAm (A.14)

Am = Mass of pesticide application [g]

Area = Applied area [m2]

RN/T = Non-target versus target ratio [-]

FTR = Fraction of chemical available for transfer [-]

DR = Fraction of residue that dissipates daily (decay rate) [1/day]

For the dermal exposure pathway via body the same equations as above exist. The dermal

exposure via body corrected for washing effect at the ith day is estimated as:

×−=

aidBodyi

aidBodywidBody EBW

EE

,,

,,,, )1(

(A.15)

Where, BWi = Body washing removal efficiency at the ith day [-]

PWB = Probability of washing body in a day [-]

EdBody,i,a = Adjusted dermal exposure via body at the ith day

),( ,,,, cidBodyLBaidBody EEMinEi

= (A.16) Where,

iLBE = Maximum dermal exposure via body [µg/kg]

B

BUhandFractotaliLB W

FSASALBE

i

×−××=

)1( (A.17)

LBi = Maximum dermal load via body [µg/cm2]

FBU = Fraction of body that is unclosed [-]

P<1-PWB (No Washing)

P>= 1-PWB (Washing)

144

EdBody,i,c = Dermal exposure via body corrected for the carryover

effect [µg/kg]

idBodyidBodyfcidBody EEEE ,1,,, +×= − (A.18) Where, EdBody,i-1 = Dermal exposure via body at the time step i-1 [µg/kg]

EdBody,i = Dermal exposure via body at the time step i [µg/kg]

B

iettnonBodyettnon

B

iettBodyiettidBody W

EDTCSRW

EDTCSRE ,argarg,arg,arg

,−− ××

+××

= (A.19)

Where, SRtarget ,t = Residue at target surfaces at the ith day (measured or

modeled) [µg/cm2]

SRnon-target, i = Residue at non-target surfaces at the ith day [µg/cm2]

TCBody = Body dermal transfer coefficient measured [cm2/hr]

EDtarget,, i = Duration of exposure to target site at the ith day [hr]

EDtarget,, i = Duration of exposure to non-target site at the ith day [hr]

Where the surface residue at the ith day is calculated using Equation (11). The average dermal

exposure from hand and body is estimated as:

m

EEE

m

iwidBody

m

ifinalidHand

Dermal

+

=∑∑

== 1,,

1,,

(A.20)

Non-dietary Ingestion Pathway

Based on the Alion suggestion, exposure via non-dietary ingestion pathway (EIngestion,i) is

a fraction of the dermal exposure via hands as:

)(,,, iwidHandiIngestion HFEE ×= (A.21)

The average exposure from non-dietary ingestion pathway is estimated as:

145

m

EE

m

iiIngestion

Ingestion

∑== 1

,

(A.22)

A.2 Probability Distributions of Model Inputs

The model input distribution assumptions are made mainly based upon SHEDS-Pesticide

model input parameters. Detailed information on the proposed model input assumptions is shown

in Table A.1. The information includes assumed probability distribution for each input,

corresponding parameters, and the sampling strategy.

146

Table A.1. Input Assumptions for the Simplified SHEDS model

Inputs Distribution Para_1a Para_2b Para_3c Unit Comment Sampling Strategy

Inhalation Pathway

Cair Triangular 0.1 1 4 [µg/m3] Concentration of applied pesticides monthly

Cb Lognormal 6.8×10-4 1.87 [µg/m3] Background air concentration

Cb∈[0, 0.85] monthly

Dinh,i Normal 550 140 [min] Inhalation exposure duration

Dinh, i∈[0, 1440] daily

PAI Normal 1.75 0.2 [-] Physical activity index

PAI∈[1, 4] daily

k Triangular 0.05 0.13 0.4 [1/day] Decay rate monthly

WB Lognormal 30 1.3 [Kg] Body weight

WB∈[16, 55] monthly

Dermal Pathway

AM Uniform 0.5 2.0 [g] Mass of Application monthly

Area Triangular 20 40 90 [m2] Applied Area monthly

BW Uniform 0.7 1.0 [-] Body washing removal efficiency daily

DR Triangular 0.1 0.2 0.4 [-] Fraction of residue that dissipates

dailymonthly

Ef Uniform 0.3 0.5 [-] Fraction of exposure from prior day daily

147

EDtarget Normal 6.0 1.4 [hr] Exposure duration at target area

EDtarget∈[0, 24] daily

EDnon-target Normal 7.5 1.6 [hr] Exposure duration at non-target area

EDnon-target∈[0, 24] daily

FTR Uniform 0.0 0.08 [-] Fraction of chemical available for

transfer daily

FBU Triangular 0.1 0.25 0.60 [-] Fraction of body that is unclosed daily

H Triangular 100 125 150 [cm] Height monthly

HW Uniform 0.3 0.9 [-] Hand washing removal efficiency daily

LH Uniform 0.1 0.7 [µg/cm2] Maximum dermal load via hand monthly

LB Uniform 0.1 0.7 [µg/cm2] Maximum dermal load via body monthly

PWB Triangular 0.3 1.0 1.0 [−] Probability of washing body monthly

PWH Triangular 0.3 1.0 1.0 [−] Probability of washing hands monthly

RN/T Uniform 0.05 0.15 [-] Non-target vs target ratio monthly

SAhandFrac Point 0.05 [-] Fraction of body surface area

associated with handsNA

TChand Lognormal 2029 1.3 [cm2/hr] Transfer coefficient (hand)

TChand∈[0, 6000] daily

TCbody Lognormal 5900 1.3 [cm2/hr] Transfer coefficient (body)

TCbody∈[0, 18000] daily

148

Non-Dietary Pathway

HF Triangular 0.025 0.05 0.25 [-] Fraction of Hands Going into Mouth

daily

Note: a. : For normal distribution, para_1 is the mean; for lognormal distribution, para_1 is geometric mean; for uniform and triangle distributions, the para_1 is minimum. b. : For normal distribution, para_2 is the standard deviation; for lognormal distribution, para_2 is geometric standard

deviation; for uniform distribution, para_2 is maximum, and for triangle distribution, the para_2 is peak. c. : For triangle distributions, the para_2 is maximum.

149

APPENDIX B

B.1 Algorithm for Application of Fourier Amplitude Sensitivity Test

Step 1: Assigning Frequencies to Inputs

• For time step t=1, assign incommensurate frequencies to inputs using the following

equation, where n represents the number of inputs and Ωn and dn are tabulated in Table

B-1:

+=Ω=

−+− inii

n

d 11

1

ωωω

(B-1)

• For example, for a case with four inputs the frequency set includes 5, 11, 19, and 23 for

ω1 to ω4, respectively.

• For time steps t=2,…,tn, where tn is the last day of the model simulation and for inputs

with daily sampling frequency generate a set of incommensurate frequencies using

Equation B-1 (See Matlab code provided in Figure B-1)

• Sample without replacement from generated set of frequencies for inputs with daily

sampling frequency (See Matlab code provided in Figure B-1) and assign these values to

daily inputs

Step 2: Setting the Search Variable

• Set the minimum sample size to be used in FAST as:

12 max +××= ωMN s (B-2)

where,

M = The interference factor (usually 4 or higher)

150

ωmax = The largest frequency among the set of ωi frequencies (including

frequencies for monthly and daily inputs) assigned to inputs in Step 1:

• Assign the sequence of sampled values for each input as:

s

sk N

Nks

′−′−

×=)12(

2π k = 1,2,…, '

sN (B-3)

where, 2

1' += s

sN

N

Step 3: Selecting Search Curves for Inputs

• Transform each input to a harmonic variable using:

)][sin( sGx iii ω= i = 1, 2, …, n (B-4)

where,

n = Number of inputs

ωi = Frequency assigned to input i in Step 3

Gi = Transformation function; )21)arcsin(1()( 1 +×= − uFuG

iPi π

F-1 = Inverse cumulative distribution function (ICDF) for the probability

distribution of inputs

• Figures B-2 and B-3 show Matlab codes for calculating the transformation functions for

selected probability distributions.

Step 4: Calculating the Main Effect Associated with Each Input

• Generate the total of 'sN Samples for each input d based on Step 3

• Feed sampled values to the model and calculate corresponding output values

• For the selected output, three alternative exposure scenarios can be considered:

o Daily exposure

151

o Incremental exposure:

>−===

−×××

=××

1,1,

)1(1

)(1

)(1

)1(1

)(1

tyyYtyY

tn

tn

tn

tn

tn

o Cumulative exposure:

∑=

×× =t

T

Tn

tn yY

1

)(1

)(1

where, n is the number of variability iterations (i.e., N’s), and yn×1(t) is selected output of

interest at time step t.

• Calculate Fourier coefficients Aj and Bj as:

[ ]

×++= ∑=

−+

0

)cos()()()(1'

1' 000

qN

jsYsYsYNA

s

N

qqNqNN

sj

q π (B-5)

[ ]

×−= ∑=

−+

0

)sin()()(1'

1' 00

qN

jsYsYNB

s

N

qqNqN

sj

q π (B-6)

Where, 2

1' −= s

qN

N and2

1'

0+

= sNN . In Equations (B-5) and (B-6),

Y(x1(s),x2(s),…,xn(s)) is shown as Y(s) for simplicity.

• Calculate the output variance as:

∑∞

=

+=1

222 )(2j

jj BAσ (B-7)

• Figures B-4 and B-5 show Matlab codes for estimating the output variance and Fourier

coefficients for different inputs, respectively

• Estimate the partial variance associated with the ith input as:

If j even If j odd

If j odd If j even

152

∑=

+=max

1

222 )(2p

ppp iii

BA ωωωσ (B-8)

• Estimate the main effect associated with input xi as:

2

2

σσ ω

ωi

iS = (B-9)

Step 5: Calculating the Total Effect Associated with Each Input

• In order to estimate the total sensitivity index for input xi, assign a new set of frequencies

to inputs based on the code provided in Figure B-6.

• Generate the total of 'sN Samples for each input d based on Step 4

• Feed sampled values to the model and calculate corresponding output values

• For the selected output, three alternative exposure scenarios can be considered:

o Daily exposure

o Incremental exposure:

>−===

−×××

=××

1,1,

)1(1

)(1

)(1

)1(1

)(1

tyyYtyY

tn

tn

tn

tn

tn

o Cumulative exposure:

∑=

×× =t

T

Tn

tn yY

1

)(1

)(1

where, n is the number of variability iterations (i.e., N’s), and yn×1(t) is selected output of

interest at time step t.

• Estimate the output variance using the code provided in Figure B-4

• Add all the spectral components in the frequency range [1, ωi/2], where the spectrum of

the Fourier series expansion is defined as 22jjj BA +=Λ . The summation of all the

153

spectral components in that frequency range provides the value of 2)( i−ωσ that is the

portion of the output variance arising from the uncertainty of all inputs except xi.

• Estimate the total sensitivity index for input xi as:

2

2)(1

σσ ω

ωi

iST −−= (B-10)

where, 2σ is the output variance based on the new frequencies.

154

Table B-1. Parameters Used in Calculating Frequency Sets Free of Interference to Fourth Order N Ωn dn N Ωn dn

1 0 4 26 385 416 2 3 8 27 157 106 3 1 6 28 215 208 4 5 10 29 449 328 5 11 20 30 163 198 6 1 22 31 337 382 7 17 32 32 253 88 8 23 40 33 375 348 9 19 38 34 441 186 10 25 26 35 673 140 11 41 56 36 773 170 12 31 62 37 875 284 13 23 46 38 873 568 14 87 76 39 587 302 15 67 96 40 849 438 16 73 60 41 623 410 17 85 86 42 637 248 18 143 126 43 891 448 19 149 134 44 943 388 20 99 112 45 1171 596 21 119 92 46 1225 216 22 237 128 47 1335 100 23 267 154 48 1725 488 24 283 196 49 1663 166 25 151 34 50 2019 0

155

Figure B-1: Matlab Code for Assigning Frequencies to Daily Inputs.

156

Figure B-2: Matlab Code for Selecting a Search Curve for Inputs with Uniform, Loguniform, and Normal Distributions.

157

Figure B-3: Matlab Code for Selecting a Search Curve for Inputs with Lognormal Distribution.

158

Figure B-4: Matlab Code for Estimating the Output Variance.

159

Figure B-5: Matlab Code for Estimating the Fourier Coefficients.

160

Figure B-6: Matlab Code for Assigning New Frequencies to Daily Inputs Used in Estimation of

Total Sensitivity Indices.

161

B.2 Algorithm for Application of Sobol’s Method

Figure B-7 shows an algorithm for application of Sobol’s method. Each step is explained in the

following.

Step 1: Generating Two Sets of Random Samples from Probability Distributions of Inputs

• Two set of random samples from probability distributions of inputs should be generated

and stored (A1, B1, A’1, and B’2, where A and B are matrices that hold the generated

values for monthly and daily inputs, respectively).

Step 2: Running the Model for Sampled Values of Inputs

• Run the model based on random samples generated in Step 1 and estimate values of

selected outputs and store the output values in a matrix format (y and y’, where y is the

matrix that holds the output values).

• For the selected output, three alternative exposure scenarios can be considered:

o Daily exposure

o Incremental exposure:

>−===

−×××

=××

1,1,

)1(1

)(1

)(1

)1(1

)(1

tyyYtyY

tn

tn

tn

tn

tn

o Cumulative exposure:

∑=

×× =t

T

Tn

tn yY

1

)(1

)(1

where, n is the number of variability iterations, and yn×1(t) is selected output of interest at time

step t.

162

Step 3: Forming New Matrices Required for Sensitivity Indices

• Form matrix Mj and Nj for input xj based on the values stored in matrices A1, B1, A’1,

and B’2 in Step 1 (see Figures B-8 and B-9)

• Figure B-10 shows a Matlab code provided for forming Mj and Nj matrices. For this

example, inputs have monthly and daily resampling frequencies.

Step 4: Estimating the Sensitivity Indices and Performing Bootstrap Simulation to

Calculated the Confidence Intervals

• Run the model based on the matrices formed in Step 3 and save the output values (YM

and YN for output values based on M and N, respectively).

• For the selected output, three alternative exposure scenarios can be considered:

o Daily exposure

o Incremental exposure:

>−===

−×××

=××

1,1,

)1(1

)(1

)(1

)1(1

)(1

tyyYtyY

tn

tn

tn

tn

tn

o Cumulative exposure:

∑=

×× =t

T

Tn

tn yY

1

)(1

)(1

where, n is the number of variability iterations, and yn×1(t) is selected output of interest at time

step t.

• Start the bootstrap simulation

o Suggested number of bootstrap simulations is 10,000 based on Archer et al. (1997)

o At each bootstrap simulation randomly sample the “output” values with replacement

and form the following new output matrices:

163

BY : Output matrix based on bootstrap sampling of Y from Step 2

'BY : Output matrix based on bootstrap sampling of Y’ from Step 2

MBY : Output matrix based on bootstrap sampling of YM

NBY : Output matrix based on bootstrap sampling of YN

o The Key point in the bootstrap sampling step is to use the same set of random

numbers when generating BY , 'BY , M

BY , and NBY .

o Main effect for input xj can be estimated using the following equation:

)())(( 2

yVyEU

S jj

−= (B-11)

where,

E2(y) = Square of expected value for the selected output:

∑=

′′′××=n

rrkrrrkrr xxxfxxxf

nyE

12121

2 ),,,(),,,(1)( LL (B-11-1)

V(y) = Variance of the selected output estimated based on BY

Uj = ∑=

+− ′′′′′××−

n

rrkjrrjjrrrrkrr xxxxxxfxxxf

n 1)1()1(2121 ),,,,,,,(),,,(

11

LLL

(B-11-2)

f(.) = Mean model output over the selected time period (e.g., one

month) for rth individual; r = 1,…, n

n = Number of variability iterations

x = Sampled values of inputs from Step 1

x’ = Resample values of inputs from Step 3

1 2

1 3

164

(1) = BY

(2) = 'BY

(3) = MBY

o Total effect for input xi can be estimated using the following equation:

)())((

12

yVyETU

S jTj

−−= − (B-12)

where,

ET2(y) = Square of expected value for the selected output:

2

121

2 ),,,(1)(

×= ∑=

n

rrkrr xxxf

nyET L (B-12-1)

U-j = ∑=

+− ′××−

n

rrkjrrjjrrrrkrr xxxxxxfxxxf

n 1)1()1(2121 ),,,,,,,(),,,(

11

LLL

(B-12-1)

f(.) = Mean model output over the selected time period (e.g., one

month) for rth individual; r = 1,…, n

n = Number of variability iterations

x = Sampled values of inputs from Step 1

x’ = Resample values of inputs from Step 3

(1) = BY

(4) = NBY

1 4

1

165

Step 5: Summarizing the Results

• Step 4 should be repeated for time steps t=1 to 30 (for one month simulation time)

• Estimate the average main and total effects at each day

• Estimate the 95% confidence intervals for the main and total effects at each day

166

Figure B-7. Flow Diagram for Application of the Sobol’s Method.

167

Figure B-8. Forming Mj and Nj Matrices for Input xj with Monthly Sampling Frequency.

168

Figure B-9. Forming Mj and Nj Matrices for Input xj with Daily Sampling Frequency.

169

Figure B-10. Matlab Code for Forming M and N Matrices for Inputs with Monthly and Daily

Sampling Frequencies.