DRAFT EPA/600/R-05/043A DO NOT CITE OR QUOTE June 2005 External Review Draft Approaches for the Application of Physiologically Based Pharmacokinetic (PBPK) Models and Supporting Data in Risk Assessment NOTICE THIS DOCUMENT IS A PRELIMINARY DRAFT. It has not been formally released by the U.S. Environmental Protection Agency and should not at this stage be construed to represent Agency policy. It is being circulated for comment on its technical accuracy and policy implications. National Center for Environmental Assessment Office of Research and Development U.S. Environmental Protection Agency Washington DC ------- DISCLAIMER This document is a draft for review purposes only and does not constitute U.S. Environmental Protection Agency policy. Mention of trade names or commercial products does not constitute endorsement or recommendation for use. 6/29/05 ii DRAFT—DO NOT CITE OR QUOTE ------- CONTENTS LIST OF BOXES v LIST OF TABLES vi LIST OF FIGURES vii LIST OF ABBREVIATIONS AND ACRONYMS viii PREFACE ix AUTHORS AND REVIEWERS x EXECUTIVE SUMMARY xi 1. INTRODUCTION 1-1 1.1. SCOPE OF DOCUMENT 1-1 1.2. INTENDED AUDIENCE 1-1 1.3. ORGANIZATION OF THE DOCUMENT 1-2 2. PHARMACOKINETIC DATA AND MODEL NEEDS IN RISK ASSESSMENT 2-1 2.1. PHARMACOKINETICS AND DOSIMETRY MODELING 2-1 2.2. DOSE-RESPONSE AND MEASURES OF DELIVERED DOSE 2-2 2.3. PHARMACOKINETIC DATA NEEDS IN RISK ASSESSMENTS 2-4 2.4. PHARMACOKINETIC MODELS IN RISK ASSESSMENT 2-7 2.5. ROLE OF PBPK MODELS IN RfC DERIVATION 2-8 2.5.1. Reference Concentration 2-8 2.5.2. Point of Departure 2-8 2.5.3. Duration Adjustment 2-9 2.5.4. Dosimetric Adjustment Factor 2-10 2.5.5. Uncertainty Factors: Role of PBPK Models 2-11 2.5.6. Summary 2-12 2.6. ROLE OF PBPK MODELS IN RfD DERIVATION 2-12 2.6.1. Reference Dose 2-12 2.6.2. Point of Departure 2-13 2.6.3. Uncertainty Factors ; 2-14 2.6.4. Summary 2-15 2.7. ROLE OF PBPK MODELS IN CANCER RISK ASSESSMENT 2-15 2.7.1. High-Dose to Low-Dose Extrapolation 2-15 2.7.2. Interspecies Extrapolation 2-16 2.7.3. Route-to-Route Extrapolation 2-17 2.7.4. Intraspecies Variability 2-17 2.7.5. Summary 2-18 2.8. USE OF PHARMACOKINETIC DATA AND MODELS IN EXPOSURE ASSESSMENT 2-18 2.8.1. Conventional Approaches 2-18 2.8.2. Role of PBPK Models 2-19 2.8.3. Summary 2-19 2.9. PHARMACOKINETIC DATA NEEDS IN RISK ASSESSMENT: SUMMARY 2-19 6/29/05 iii DRAFT—DO NOT CITE OR QUOTE ------- CONTENTS (continued) 3. EVALUATION OF PBPK MODELS INTENDED FOR USE IN RISK ASSESSMENT 3-1 3.1. MODEL PURPOSE 3-1 3.2. MODEL STRUCTURE 3-2 3.3. MATHEMATICAL REPRESENTATION 3-3 3.4. PARAMETER ESTIMATION 3-7 3.4.1. Physiological Parameters 3-8 3.4.2. Partition Coefficients 3-11 3.4.3. Biochemical Parameters 3-12 3.5. COMPUTER IMPLEMENTATION 3-13 3.6. MODEL EVALUATION 3-16 3.7. SENSITIVITY, UNCERTAINTY, AND VARIABILITY ANALYSES 3-21 3.7.1. Sensitivity Analysis 3-21 3.7.2. Variability Analysis 3-23 3.7.3. Uncertainty Analysis 3-25 3.8. DEVELOPING PBPK MODELS FOR USE IN RISK ASSESSMENT: STRATEGIES FOR DEALING WITH DATA-POOR SITUATIONS 3-27 3.8.1. Minimal Data Needs for Constructing PBPK Models 3-27 3.8.2. Surrogate Data for Interspecies and Interchemical Extrapolations 3-28 3.9. EVALUATION OF PBPK MODELS: SUMMARY 3-30 4. APPLICATION OF PBPK MODELS IN RISK ASSESSMENT 4-1 4.1. CHOOSING PBPK MODELS APPROPRIATE FOR USE IN RISK ASSESSMENT 4-1 4.2. EVALUATION OF DOSE METRICS FOR PBPK MODEL-BASED ASSESSMENTS 4-4 4.3. EXAMPLES OF THE USE OF PBPK MODELS IN RISK ASSESSMENT .... 4-9 4.3.1. High-Dose to Low-Dose Extrapolation 4-9 4.3.2. Interspecies Extrapolation 4-10 4.3.3. Route-to-Route Extrapolation 4-12 4.3.4. Duration Adjustment 4-14 4.3.5. Intraspecies Extrapolation 4-15 4.3.6. RfD Derivation 4-16 4.3.7. RfC Derivation 4-17 4.3.8. Cancer Risk Assessment (unit risk estimates, RfC, and RfD) 4-18 4.3.9. Mixture Risk Assessment 4-21 4.4. LINKAGE TO PHARMACODYNAMIC MODELS 4-22 GLOSSARY G-l REFERENCES R-l 6/29/05 iv DRAFT—DO NOT CITE OR QUOTE ------- LIST OF BOXES Box 2-1. Examples of dose metrics useful for exploring dose-response relationships 2-4 Box 2-2. Role of PBPK models in the RfC process 2-12 Box 2-3. Role of PBPK models in the RfD process 2-15 Box 2-4. Role of PBPK models in cancer risk assessment 2-18 Box 2-5. Role of PBPK models in exposure assessment 2-19 Box 4-1. Questions and issues to be considered while evaluating the adequacy of a PBPK model 4-3 6/29/05 v DRAFT—DO NOT CITE OR QUOTE ------- LIST OF TABLES Table 3-1. Equations of a four-compartment PBPK model to simulate the inhalation exposure of volatile organic compounds 3-5 Table 3-2. Equations used for describing diffusion-limited uptake in PBPK models 3-6 Table 3-3. Reference physiological parameters for mice, rats, and humans 3-9 Table 3-4. Range of values of the volume and perfusion of select tissues in the mouse 3-9 Table 3-5. Range of values of the volume and perfusion of select tissues in the rat 3-10 Table 3-6. Range of values of perfusion of select tissues in humans 3-10 Table 3-7. Examples of simulation software used for PBPK modeling 3-14 Table 4-1. Dose metrics used in PBPK model-based cancer and noncancer risk assessments 4-7 Table 4-2. Relationship between tumor prevalence and dichloromethane metabolites produced by microsomal and glutathione pathway for the bioassay conditions (methylene chloride-dose response in female mice) 4-20 Table 4-3. Examples of biologically based models of endpoints and processes of toxicological relevance 4-25 6/29/05 VI DRAFT—DO NOT CITE OR QUOTE ------- LIST OF FIGURES Figure 2-1. Relationship between the exposure concentration and adverse response for a hypothetical chemical 2-3 Figure 2-2. Rat-human extrapolation of exposure concentrations of toluene on the basis of equivalent dose metrics (AUC of toluene in blood; 3.8 mg/L-hr) 2-6 Figure 3-1. Conceptual representations of PBPK models for (A) toluene, (B) ethylene oxide, (C) 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD), and (D) vinyl acetate. 3-4 Figure 3-2. Comparison of PBPK model simulations (solid lines) with experimental data (symbols) 3-18 Figure 3-3. Sensitivity ratios associated with certain input parameters of a PBPK model.. 3-22 Figure 3-4. Illustration of the use of the Monte Carlo approach for simulating the distribution of internal concentration versus time on the basis of population distributions of PBPK model parameters 3-25 Figure 4-1. Decision tree for selecting PBPK models appropriate for use in risk assessment 4-2 Figure 4-2. Examples of measure of tissue exposure to toxic moiety for risk assessment applications 4-6 Figure 4-3. High-dose to low-dose extrapolation of dose metrics using PBPK model for toluene 4-11 Figure 4-4. Oral-to-inhalation extrapolation of the pharmacokinetics of chloroform on the basis of same AUC in blood (7.06 mg/L-hr) 4-13 Figure 4-5. Duration adjustment (4 hr to 24 hr) of toluene exposures in rats, based on equivalent AUC (2.4 mg/L-hr) 4-14 Figure 4-6. Estimation of the interindividual factor from the 50th (median) and 95th percentile values of a dose metric simulated with a probabilistic PBPK model 4-15 Figure 4-7. PBPK model predictions of glutathi one-path way metabolites in liver in mice 4-20 Figure 4-8. Relationship between the dose metric ((irnol metabolized/g liver/hr) simulated by PBPK model and the cell killing inferred from pharmacodynamic model for chloroform 4-23 6/29/05 VII DRAFT—DO NOT CITE OR QUOTE ------- LIST OF ABBREVIATIONS AND ACRONYMS ADME absorption, distribution, metabolism, and excretion AUC area under the curve BBDR biologically based dose-response (model) BMD benchmark dose BMC benchmark concentration Cmax maximal concentration CSF cancer slope factor DAF dosimetric adjustment factor HEC human equivalent concentration HI hazard index IUF interspecies uncertainty factor IUR inhalation unit risk IVF intraspecies variability factor LOAEL lowest-observed-adverse-effect level NOAEL no-observed-adverse-effect level MBDE mass balance differential equation MO A mode of action MVK Moolgavkar-Venzon-Knudson OSF oral slope factor PK pharmacokinetic PBPK physiologically based pharmacokinetic POD point of departure QSAR quantitative structure-activity relationship RfD reference dose RfC reference concentration 6/29/05 vui DRAFT—DO NOT CITE OR QUOTE ------- PREFACE This draft report, Approaches for the Application of Physiologically Based Pharmacokinetic (PBPK) Models and Supporting Data in Risk Assessment, addresses the application and evaluation of PBPK models for risk assessment purposes. These models represent an important class of dosimetry models that are useful for predicting internal dose at target organs for risk assessment applications. This report is primarily meant to serve as a learning tool for EPA scientists and risk assessors who may be less familiar with the field. In addition, this report can be informative to PBPK modelers within and outside the Agency, as it provides an assessment of the types of data and models that the EPA requires for consideration of a model for use in risk assessment. This draft report will be externally peer-reviewed by an independent panel of experts prior to making the report final. 6/29/05 ix DRAFT—DO NOT CITE OR QUOTE ------- AUTHORS AND REVIEWERS The National Center for Environmental Assessment, Office of Research and Development, was responsible for the preparation of this document. It was developed under EPA Contract No. 4W-0322-NASX. EPA PROJECT OFFICERS Femi Adeshina National Center for Environmental Assessment Office of Research and Development U.S. Environmental Protection Agency Washington, DC Chad wick Thompson National Center for Environmental Assessment Office of Research and Development U.S. Environmental Protection Agency Washington, DC AUTHORS Kannan Krishnan, University of Montreal, Alberta, Canada Hugh Barton, NHEERL-RTP Rob DeWoskin, NCEA-RTP Bob Sonawane, NCEA-W Chadwick Thompson, NCEA-W INTERNAL EPA REVIEWERS Jerry Blancato, NERL-LV Weihsueh Chiu, NCEA-W Joyce Donohue, OW Hisham El-Masri, NCEA-W Marina Evans, NHEERL-RTP Lynn Flowers, NCEA-IRIS Karen Hammerstrom, NCEA-W John Lipscomb, NCEA-Cin Allen Marcus, NCEA-IRIS Dierdre Murphy, OAQPS Alberto Protzel, OPP-W Paul Schlosser, NCEA-RTP Woodrow Setzer, NHEERL-RTP 6/29/05 x DRAFT—DO NOT CITE OR QUOTE ------- EXECUTIVE SUMMARY Physiologically based pharmacokinetic (PBPK) models represent an important class of dosimetry models that are useful for predicting internal dose at target organs for risk assessment applications. Approaches for the Application of Physiologically Based Pharmacokinetic (PBPK) Models and Supporting Data in Risk Assessment addresses the following questions: Why are risk assessors interested in using PBPK models? How are PBPK models evaluated for use in a risk assessment? What are the questions or data gaps in a risk assessment that can be addressed by PBPK models? Because PBPK modeling efforts to date have largely focused on chemicals that cause systemic effects (i.e., water-insoluble gases and some nonvolatile organics), this document draws upon the experience and literature on these efforts. The text is organized into four chapters. Chapter 1 outlines the scope of the document, the intended audience, and the topics covered in the remaining chapters. Chapter 2 presents the rationale for using PBPK models in risk assessment and the pharmacokinetic data and models needed to derive a reference dose (RfD), a reference concentration (RfC), and unit risk estimates in cancer risk assessment (e.g., cancer slope factor). Chapter 3 describes how models are evaluated, the main model characteristics to review, and the on-going development of acceptance criteria for model use in risk assessment. Chapter 4 discusses applications of PBPK model simulations within the current U.S. Environmental Protection Agency risk assessment framework. Useful resources are provided as appendices, including (1) a list of publications relating to PBPK modeling and its use in health risk assessment, (2) a compilation of parameters of several published PBPK models, and (3) a list of published algorithms for estimating parameters for PBPK models of environmental chemicals. PBPK models consist of a series of equations for compartments, fluid flows, and chemical reactions that represent real biological tissues and physiological processes in the body and that simulate the absorption, distribution, metabolism, and excretion of chemicals that enter the body. PBPK models are designed to estimate an internal dose of a proposed toxic moiety to a target tissue(s) or some appropriate surrogate dose metric for a target tissue dose. The choice of an internal dose metric is based on an understanding of the chemical's mode of action. The internal dose metric (sometimes called the biologically effective dose) replaces the applied dose in a quantitative dose-response assessment, with the intent of reducing the uncertainty inherent in 6/29/05 xi DRAFT—DO NOT CITE OR QUOTE ------- using the applied dose to derive risk values. This reduction in uncertainty and improved scientific basis for the risk value are the main advantages of PBPK models and the reasons for the growing interest in their use. PBPK models also can simulate an internal dose from exposure conditions of interest where no data are available, i.e., to extrapolate to conditions beyond those of the data set used to develop the model. An important and active area of research is the characterization of the uncertainty in risk assessments based on PBPK model results compared with the uncertainty in results based on the applied dose. Examples of PBPK model applications in risk assessments include interspecies extrapolation of the dose-response (based on estimates of the internal dose), route-to-route extrapolation, estimation of response from varying exposure condition, estimation of human variability (within the whole population or subpopulations), and high-to-low dose extrapolation. PBPK models used in risk assessments should, at a minimum, (1) contain a compartment that is either identified with the target tissue, contains the target tissue, or is identified as a surrogate for the target tissue; (2) have defensible physiological parameter values that are within the known plausible range; and (3) have undergone a thorough evaluation for their structure, implementation, and predictive capability. As a resource, appendices to this document provide a compilation of parameters of several published PBPK models, a list of available animal- alternative algorithms for parameter estimation, and key references. Evaluation of PBPK models intended for risk assessments includes a review of the model purpose, model structure, mathematical representation, parameter estimation (calibration), and computer implementation. Criteria for acceptance of a PBPK model for use in risk assessment include the following: (1) the model represents the species and life stage of relevance to a particular risk assessment, (2) the model has been evaluated and peer-reviewed for the adequacy of its structure and parameters, and (3) the model provides adequate simulations of the concentration of the toxic moiety (parent chemical or metabolite) in the target organ (or a surrogate compartment) following the relevant route(s) of exposure and over the time-course for which the chemical is present in that tissue. When a PBPK model is available for the appropriate test species, it is used to estimate the value of internal dose metrics, which are then use to derive a given point of departure (e.g., NOAEL, LOAEL, BMD, BMC) for use in dose-response analyses for toxicity endpoints, including cancer, chronic toxicity, and other toxicity endpoints. Some risk assessment applications can be accomplished using only a model for the test species, e.g., prediction of the 6/29/05 xii DRAFT—DO NOT CITE OR QUOTE ------- toxicity in that species by another route of exposure for purposes of route extrapolation. For most applications, a version of the PBPK model is also developed to simulate kinetics in humans for estimating the applied dose to a human needed to achieve the relevant internal dose predicted from the animal toxicity or, less frequently, human clinical or epidemiology study. PBPK model analysis is accepted as a scientifically sound approach to estimating the internal dose of a chemical at a target site and as a means to evaluate and describe the uncertainty in risk assessments. 6/29/05 xiii DRAFT—DO NOT CITE OR QUOTE ------- 1 1. INTRODUCTION 2 3 1.1. SCOPE OF DOCUMENT 4 The objective of this document is to provide a description of approaches for using 5 physiologically based pharmacokinetic (PBPK) data and models in human health risk 6 assessment. The document focuses primarily on the evaluation and use of PBPK models for 7 predicting internal dose at target organs in risk assessment applications. Many of the past efforts 8 on PBPK modeling have focused on water-insoluble gases that cause systemic toxicity (i.e., 9 producing effects remote from the site of exposure), as well as some nonvolatile organics. This 10 document primarily draws on the experience and literature on these efforts. The approaches are 11 not intended, for example, to apply directly to agents, gaseous or particulate, where the target 12 organ is the respiratory tract; the reference concentration (RfC) methodology (U.S. EPA, 1994) 13 should be consulted for these circumstances. However, the elements of discussion found in this 14 document are also conceptually applicable, in a broader sense, to other kinds of dosimetry 15 models and a wider range of substances. 16 In developing this document, it was assumed that risk assessors are familiar with the 17 basic concepts of PBPK modeling and that model developers are familiar with the basic concepts 18 of risk assessment; therefore, the document contains only brief descriptions of PBPK modeling 19 and risk assessment methods. Appropriate references to secondary review articles and reports 20 from which readers can obtain additional information are provided. 21 Finally, it is important to realize that the application of PBPK models in risk assessment 22 is evolving. Thus, this document does not specify (or recommend) when the effort to construct 23 and apply PBPK modeling is justified by the additional insight it provides; rather it highlights 24 some of the benefits of PBPK modeling in risk assessment. 25 26 1.2. INTENDED AUDIENCE 27 The document was prepared with two primary audiences in mind: (1) risk assessors who 28 need to know about the potential applications of PBPK models in risk assessments, and (2) 29 PBPK model developers who need to be familiar with potential applications in health risk 30 assessment. 31 6/29/05 1-1 DRAFT—DO NOT CITE OR QUOTE ------- 1 1.3. ORGANIZATION OF THE DOCUMENT 2 Three key chapters form the main core of this document. They describe what risk 3 assessors need in terms of pharmacokinetic data, and why (Chapter 2); how to evaluate PBPK 4 models for use in risk assessments (Chapter 3); and how to use PBPK models in risk assessments 5 to address specific areas of uncertainty (Chapter 4). 6 Chapter 2 addresses data needs in terms of reference dose (RfD) and RfC derivation, as 7 well as predictive estimates in cancer risk assessment. This chapter also contains a brief 8 discussion on the minimal data requirements for constructing PBPK models, as well as the use of 9 pharmacokinetic data and PBPK models to improve exposure assessments. 10 Chapter 3 presents an approach and some criteria for evaluating PBPK models intended 11 for use in risk assessments. The issues relating to the evaluation of PBPK models that are 12 discussed in this chapter will facilitate the assessor's decision on whether or not an available 13 model is adequate and scientifically defensible for use in reducing uncertainties in a given risk 14 assessment. In this regard, the PBPK modeling issues are considered under each of the 15 following topic areas: model structure, mathematical description, parameter estimation 16 (calibration), computer implementation, and evaluation. Current criteria as well as accepted 17 methods are identified and then assembled to facilitate the identification of PBPK models that 18 meet the requirement for use in risk assessment. 19 Chapter 4 discusses how PBPK models and data can be applied within the current 20 Agency risk assessment framework to address specific areas of uncertainty. The following types 21 of PBPK model applications in risk assessment are presented in this chapter: high-dose to low- 22 dose extrapolation, interspecies extrapolation, intraspecies extrapolation, route-to-route and 23 scenario extrapolation, mixture risk assessment, and linkage with pharmacodynamic models. 24 This chapter also highlights how PBPK models are used in cancer and noncancer assessments. 25 The appendices provide a list of publications relating to PBPK modeling and its use in 26 health risk assessment, a compilation of parameters of several published PBPK models, and a list 27 of published algorithms for estimating parameters for PBPK models of environmental chemicals. 28 29 6/29/05 1 -2 DRAFT—DO NOT CITE OR QUOTE ------- 1 2. PHARMACOKINETIC DATA AND MODEL NEEDS IN RISK ASSESSMENT 2 3 2.1. PHARMACOKINETICS AND DOSIMETRY MODELING 4 Pharmacokinetics (pharmakon + kinetics; pharmakon (Greek) = drugs and poisons; 5 kinetics = change as a function of time) involves the study of the time course of the parent 6 chemical or metabolite concentrations or amounts in biological fluids, tissues, and excreta and 7 the construction of mathematical models to interpret such data (Wagner, 1981). The time course 8 of the concentration of a chemical or its metabolite in biota is determined by the rate and extent 9 of absorption, distribution, metabolism, and excretion (ADME). The pharmacokinetics or 10 ADME of a substance determines the delivered dose or the amount of chemical available for 11 interaction in the tissues. Relating adverse response observed in biota to an appropriate measure 12 of delivered dose (e.g., concentration of the toxic chemical in the target tissue) rather than 13 administered dose or exposure concentration is likely to improve the characterization of many 14 dose-response relationships. 15 A range of modeling approaches is used to characterize exposures and resulting delivered 16 doses. This variety of approaches reflects differences in chemical and physical characteristics 17 (e.g., stable or reactive gases, particulate matter, lipophilic organics, water-soluble compounds) 18 and their ability to cause contact site or systemic toxic effects (Andersen and Jarabek, 2001; + 19 Overton, 2001; U.S. EPA, 1994; 2004). Many drugs and toxic chemicals that cause systemic 20 effects are absorbed following oral exposures, so compartmental and noncompartmental 21 pharmacokinetic modeling has focused on these characteristics (O'Flaherty 1981; Renwick 22 2001). Compartmental modeling, generally described as a set of mathematical equations that 23 describe empirical pharmacokinetic data concerning a chemical, evolved to include mathematical 24 descriptions of biological processes (e.g., partition coefficients, tissue volumes), which is the 25 focus of this document. 26 Because the respiratory tract is a frequent site of both exposure and toxiciry, it has been a 27 particular focus for a range of modeling approaches. These approaches include those developed 28 for gases of various reactivity and solubility, as well as for particulate matter (see U.S. EPA, 29 1994). More refined approaches are available and useful for reactive gases and particulate 30 matter, including two-dimensional and three-dimensional modeling such as computational fluid 31 dynamics models (Kimbell et al., 1993; Martonen et al., 2001; Overton, 2001, U.S. EPA, 2004). 32 Predicting behaviors of volatile anesthetics, including compounds now used exclusively as 6/29/05 2-1 DRAFT—DO NOT CITE OR QUOTE ------- 1 industrial chemicals, was a driving force for the development of PBPK models (Krishnan and 2 Andersen, 2001). 3 Another significant factor influencing modeling approaches is the role of metabolism. 4 For example, a range of volatile compounds can cause nasal or other respiratory toxicities that 5 require metabolism of the chemical in respiratory tract tissue. These kinds of chemicals have 6 been modeled using elaborate PBPK models of the nose (with or without representation of the 7 rest of the body) that have been linked with computational fluid dynamic models to describe the 8 deposition of chemical in different regions of the nose (Frederick et al., 2002; Bogdanffy et al., 9 2001). Similarly, compounds that are relatively water soluble require specific approaches that 10 address aspects such as fractional absorption or "wash-in, wash-out" effects (Johanson, 1991; 11 Medinsky et al., 1993; Perkins et al., 1995). 12 Although the approaches detailed in the RfC methodology do not address all of these 13 aspects, there is recognition of the need for additional approaches that address these and other 14 challenging aspects of respiratory tract dosimetry (U.S. EPA, 1994). 15 The relevant modeling approach, therefore, depends on the physical and chemical 16 characteristics of the material, the method and route of exposure or delivery, and the toxicities 17 under consideration. All of these modeling approaches attempt to describe the dose delivered to 18 the relevant areas of the body, whether that is a region of the respiratory tract or skin, or systemic 19 delivery through the blood supply to target organs. These approaches permit estimation of some 20 measure of delivered dose for improved understanding of the dose-response relationship. 21 22 2.2. DOSE-RESPONSE AND MEASURES OF DELIVERED DOSE 23 Dose-response relationships that appear unclear or confusing at the administered dose 24 level can become more understandable when expressed on the basis of internal dose of the 25 chemical. Figure 2-1 depicts the case of a hypothetical chemical for which the correlation between 26 dose and response is weak or complex (Panel A). However, once the relationship is based on 27 internal dose, there emerges a clear and direct relationship between dose and response (Panels B 28 and C). The major advantage of constructing dose-response relationships on the basis of internal 29 or delivered dose is that it can provide a stronger biological basis for conducting extrapolations and 30 for comparing responses across studies, species, routes, and dose levels (Andersen et al., 1987; 31 Clewell and Andersen, 1997; Aylward et al., 1996; Benignus et al., 1998; Melnick and Kohn, 32 2000). 6/29/05 2-2 DRAFT—DO NOT CITE OR QUOTE ------- Hill. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 linear Weibull 60-| TJ O !_ 40-\ ||> 30- I 20- *o £ 10 - (0 a: 0 B 1000 2000 3000 Exposure concentration (ppm) 30 - 25 • § 20 o t 15- CL £ 10- 4000 1 10 100 1000 Exposure concentration (ppm) 10000 10 20 30 40 50 Rate of amount metabolized (mg/hr) Figure 2-1. Relationship between the exposure concentration and adverse response for a hypothetical chemical. Panel A depicts the case of a chemical for which the correlation between dose and response is weak or complex, along with equally plausible curve fits (linear, Hill, and Weibull). This dose-response relationship is improved when it is based on an appropriate measure of internal dose (Panels B and C). The use of blood and tissue concentrations for relating dose and response in exposed organisms has long been recognized in pharmacology (e.g., Wagner, 1981). The target tissue dose closely related to ensuing adverse responses is often referred to as the "dose metric" (Andersen and Dennison, 2001). Dose metrics used for risk assessment applications should reflect the biologically active form of the chemical (parent chemical, metabolites, or adducts), its level (concentration or amount), duration of internal exposure (instantaneous, daily, lifetime, or specific window of susceptibility), and intensity (peak, average, or integral), as well as the appropriate biological matrix (e.g., blood, target tissue, surrogate tissues). For assessment of health risks related to lifetime exposure of systemically acting chemicals, in the absence of mode 6/29/05 2-3 DRAFT—DO NOT CITE OR QUOTE ------- 1 2 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 of action (MO A) information to the contrary, the integrated concentration of the toxic form of chemical over lifetime (i.e., area under the concentration versus time curve [AUC]) in target tissue has been considered to be the appropriate dose metric (Collins, 1987; Voisin et al., 1990; Clewell et al., 2002). When the carcinogen is not the parent chemical but a reactive intermediate, the amount of metabolite produced per unit time or the amount of metabolite in target tissue over a period of time (e.g., mg metabolite/L tissue during 24 hr) has been used as the dose metric (Andersen and Dennison, 2001). For developmental effects, the dose surrogate is defined in the context of window of susceptibility for a particular gestational event (e.g., Welsch et al., 1995; Luecke et al., 1997). Even though the AUC and rate of metabolite formation are among the most Box 2-1. Examples of dose metrics useful for exploring dose-response relationships Parent chemical • Peak concentration • Average concentration • Amount or quantity • AUC Metabolite • Peak concentration • Average concentration • Amount or quantity • Rate of production • Cumulative rate of formed/time/L tissue • AUC Miscellaneous • Receptor occupancy (extent/duration) • Macromolecular adduct levels • Depletion of cofactors commonly investigated dose metrics, other surrogates of tissue exposure may also be appropriate for risk assessment purposes, depending on the chemical and its MOA (Clewell et al., 2002). Box 2-1 lists dose metrics that can be used to derive dose-response relationships for risk assessment. 2.3. PHARMACOKINETIC DATA NEEDS IN RISK ASSESSMENTS The dose-response assessment portion of the risk assessment process can be used to determine a point of departure (POD) for one or more of the most sensitive critical effects, based on the relationship between administered dose and observed responses in laboratory or field studies. For noncancer and nonlinear cancer assessments, adjustments to the POD are then made to account for uncertainties in the estimate of a POD for humans and to protect the most sensitive human subpopulation exposed by specific routes (e.g., inhalation, oral). This process frequently requires conducting interspecies, intraspecies, high-dose to low-dose, duration, and exposure route extrapolations of responses. Based on information about the MOA of the compound, an 6/29/05 2-4 DRAFT—DO NOT CITE OR QUOTE ------- 1 appropriate approach for the extrapolation can be identified. Such an approach implies that 2 when the value of the dose metric is identical in two situations (rat vs. human, oral vs. inhalation, 3 6-hr exposure vs. 24-hr exposure), the two administered doses are pharmacokinetically 4 equivalent For example, exposure of rats to 50 ppm toluene for 6 hr and of humans to 17.7 ppm 5 toluene for 24 hr yields the same blood AUC (3.8 mg/L-hr), implying that these exposure 6 scenarios in rats and humans are pharmacokinetically equivalent (Figure 2-2). Alternatively, if 7 one were to assume that the pharmacokinetics in humans were identical to those in rats, an 8 equivalent 24 hr exposure would be calculated to be 12.5 ppm (i.e., 6 hr x 50 ppm is equivalent 9 to24hr x 12.5 ppm). 10 For reactive gases that cause contact site toxicity, the concentration of chemical or its 1.1 delivery rate to the target may be the relevant dose metric, whereas for anesthetic effects of 12 volatile organics, the current concentration in the blood has been an effective dose metric. In 13 these cases, the acute effects are more —or entirely— dependent upon concentration than on 14 time, so the extrapolations would be dependent upon that dose metric, potentially without 15 adj ustments for duration of exposure. 16 The most robust pharrnacokinetic data set needed for risk assessment would consist of the 17 time-course data on the dose metric associated with exposure scenarios and doses used in the 18 critical studies chosen for the assessment (e.g., animal bioassays or human epidemiological 19 studies) and relevant human exposure conditions. An example of such a dose metric is the 20 concentration of a toxic metabolite in target tissue over a 24 hr period in the test species and in 21 humans. This information should be obtained for the window of susceptibility, route and 22 scenario of exposure associated with the critical study as well as for the window of susceptibility, 23 appropriate route, and exposure scenarios in humans. If such a pharmacokinetic data set is 24 available for each scenario of interest, then there is no real need for any pharmacokinetic models, 25 even though their availability can facilitate simulation of other potential scenarios of interest and 26 critical determinants of tissue dose. 27 In almost all cases, dose metric measurements associated with human exposures to 28 environmental chemicals will not be available. Even the available animal pharmacokinetic data 29 may not correspond to the active toxic moiety, relevant route, or appropriate dose levels. In the 30 absence of experimental data on the biologically active form of the chemical in target tissues, 31 data on blood concentration of the parent chemical, urinary metabolite levels, or fraction 32 6/29/05 2-5 DRAFT—DO NOT CITE OR QUOTE ------- RAT I I 100- 10- 1 - 0.1 - 0.01 - 0.001 - 0.0001 8 12 16 Time (hours) 20 24 HUMAN Venous blood concentration 10 - 1 - i °" •§ 0.01 - 0.001 - n nnm . /•"^ 0 4 8 12 16 20 24 Time (hours) 2 3 4 5 6 7 8 9 10 11 12 13 14 Figure 2-2. Rat-human extrapolation of exposure concentrations of toluene on the basis of equivalent dose metrics (AUC of toluene in blood; 3.8 mg/L- hr). Rat exposures are for 6 hr to 50 ppm; human exposures are for 24 hr to 17.7 ppm. Both exposures yield the same AUC, as determined using species-specific PBPK models published by Tardif et al. (1997). Similar exercises can be done to determine the exposure concentrations that yield equivalent peak concentrations (Cmax) in rats and humans. absorbed may be used as a surrogate of dose metrics. These data can be used to develop a PBPK model to estimate the level of the toxic moiety of interest, and the uncertainty in those estimates can be formally characterized. 6/29/05 2-6 DRAFT—DO NOT CITE OR QUOTE ------- 1 2.4. PHARMACOKINETIC MODELS IN RISK ASSESSMENT 2 Regulatory agencies have a mandate to develop guidelines for toxicity testing and 3 formulate risk values based on the current understanding of the dose-response relationship. 4 Reference values correspond to an estimate of an exposure level for a given duration that is 5 likely to be without an appreciable risk of adverse effects over the lifetime in humans, including 6 sensitive subpopulations. The reference values developed at EPA include reference 7 concentration (RfC) for chronic inhalation exposures and reference dose (RfD) for chronic oral 8 exposures. For chronic oral and inhalation cancer risk assessments with an unknown or a linear 9 MOA (e.g., mutagenic carcinogens), EPA develops unit risk estimates, such as cancer slope 10 factor (CSF) and inhalation unit risk (IUR) estimates. The underlying assumption in these 11 processes is that the exposure concentration (or applied dose) results in a toxic chemical moiety 12 in a target organ that will be less than or equal to a level not associated with significant adverse 13 responses during a lifetime (reference value) or that yields a likely risk at or below the estimated 14 lifetime risk (unit risk). 15 Even though a key factor in the induction of adverse effects is the presence of the toxic 16 form of a chemical in the target organ, it is rare that data are available on the time course of the 17 toxic moiety of chemicals in the target tissue(s) in humans. Even in animal studies, it is more 18 practical to get measures of blood, plasma, and urinary concentrations of toxic chemicals and 19 their metabolites than to get data on the time course on the actual tissue dose of a toxic moiety 20 relevant for use in risk assessment. Pharmacokinetic models are therefore used to simulate the 21 tissue concentration of toxic substances. There are both noncompartmental and compartmental 22 pharmacokinetic models (Renwick, 2001); however, we will focus on compartmental models. 23 Among the compartmental pharmacokinetic models, those that allow for prediction of 24 target organ concentration are most appropriate and useful for conducting extrapolations 25 essential for risk assessment. These models are commonly referred to as PBPK. models 26 (Himmelstein and Lutz, 1979; Rowland, 1985; Leung, 1991; Andersen, 1995; Krishnan and 27 Andersen, 2001). By simulating the kinetics and dose metric of chemicals, they attempt to 28 reduce the uncertainty related to the interspecies, intraspecies, high-dose to low-dose, route-to- 29 route, and exposure scenario extrapolations essential in the context of RfC, RfD, and cancer unit 30 risk estimate processes. The following sections discuss how the PBPK models are used in the 31 context of health risk assessment. 32 6/29/05 2-7 DRAFT—DO NOT CITE OR QUOTE ------- 1 2.5. ROLE OF PBPK MODELS IN RfC DERIVATION 2 2.5.1. Reference Concentration 3 The RfC corresponds to an estimate of continuous inhalation exposure (mg/m3) for a 4 human population, including sensitive subgroups, that is likely to be without an appreciable risk 5 of adverse health effects during a lifetime (U.S. EPA, 1994). Notationally, RfC is defined as: 6 7 RJC = POD[HEC]/UF 8 9 where: 10 POD[HEC] = POD (no-observed-adverse-effect level [NOAEL], lowest-observed-adverse- 11 effect level [LOAEL], or benchmark concentration [BMC]) dosimetrically 12 adjusted to a human equivalent concentration (HEC) 13 14 UF = uncertainty factors to account for the extrapolations associated with the POD (i.e., 15 interspecies differences in sensitivity, intraspecies extrapolation, subchronic-to- 16 chronic extrapolation, LOAEL-to-NOAEL extrapolation, and incompleteness of 17 database) 18 19 The starting point for an RfC derivation is the identification of the POD for the critical 20 effect in a key study. Subsequent steps involve (1) adjustment for the difference in duration 21 between experimental exposure (e.g., 6 hr) and expected human exposure (24 hr); (2) calculation 22 of the HEC, based on dosimetric adjustments; and (3) application of uncertainty factors. The 23 benefit of using PBPK models in the RfC process is discussed below. Specifically, the role of 24 PBPK models in determining the POD, duration adjustment factor, HEC, and uncertainty factors 25 is presented in Sections 2.5.2 through 2.5.5. 26 27 2.5.2. Point of Departure 28 Ideally, the POD used in the RfC process should be the inhalation route-specific NOAEL, 29 LOAEL, or BMC. These PODs essentially correspond to exposure concentrations of an 30 experimental or field study (NOAEL, LOAEL) or to the lower confidence limit (95th percentile) 31 of the exposure concentration (BMCL) associated with a specified response level (generally in 32 the range of 1 to 10 %; e.g., BMCLio%) derived from statistical analysis of experimental dose- 33 response data (U.S. EPA, 1994, 2000a). 34 Route-to-route extrapolation can be conducted on the basis of equivalent potential doses 35 when information on the POD was available only for a noninhalation route of exposure (e.g., oral 6/29/05 2-8 DRAFT—DO NOT CITE OR QUOTE ------- 1 route) (Pauluhn, 2003). For example, the NOAEL (mg/kg/day) associated with a particular 2 exposure route (e.g., oral) was converted to milligrams per day and then to equivalent inhaled 3 concentration, based on breathing rate and body weight of the test species. This simplistic 4 approach assumes that the rates of absorption, distribution, metabolism, excretion, and tissue 5 dosimetry of chemicals are the same for a given total dose, regardless of the exposure route and 6 intake rate. Such an approach essentially neglects the route-specific differences in 7 pharmacokinetics—more specifically, absorption, clearance, and first-pass effects. When data 8 on the route-specific fraction absorbed are available, they are used to determine the equivalent 9 inhalation concentration on the basis of equivalent absorbed doses (U.S. EPA, 1999a), but 10 preference would be given to more complete pharmacokinetic modeling. 11 The POD for RfC derivation cannot be identified or established with only 12 pharmacokinetic data or PBPK models, independent of the dose-response data. On the other 13 hand, better characterization of dose-response relationships has been demonstrated by the use of 14 integrated pharmacokinetic-pharmacodynamic models (e.g., Gearhart et al., 1990,1994; 15 Timchalk et al., 2002). Such models might be capable of predicting response and thus 16 determining a POD if sufficient data are available. More commonly, however, PBPK modeling 17 will be useful for conducting route-to-route extrapolation on the basis of equivalent delivered 18 dose from PODs identified from NOAEL, LOAEL, or BMC (e.g., oral to inhalation). 19 20 2.5.3. Duration Adjustment 21 RfCs are intended for continuous exposure of human populations, such that the POD used 22 in its derivation should correspond to 24 hr/day exposures (U.S. EPA, 1994). Because the PODs 23 are frequently obtained from animal exposures or occupational exposures that occur for 6 to 8 24 hr/d, 5 d/wk, adjustment to a continuous 24 hr exposure is conducted on the basis of hours per 25 day and days per week (i.e., 6/24 x 5/7), which essentially results in a lower concentration for 26 continuous exposures (U.S. EPA, 2002). This simple adjustment, generally referred to as 27 Haber's Rule, implies that the AUC (i.e., concentration x time [C x t]) and not peak 28 concentration is the dose metric associated with the toxicity endpoints (U.S. EPA, 2002). When 29 data indicate that a given toxicity is more dependent on concentration than on duration (time), 30 this adjustment would not be used. If neither dose metric is demonstrable experimentally to be 31 the appropriate measure of internal dose, the Agency uses adjustment to a continuous inhalation 6/29/05 2-9 DRAFT—DO NOT CITE OR QUOTE ------- 1 exposure based on the C x t relationship as a matter of health-protective policy (U.S. EPA, 2 2002). 3 Depending on the dose metric identified or hypothesized to be the most appropriate for 4 the chemical and endpoint, the duration-adjusted exposure values can be obtained with PBPK 5 models (Jarabek, 1994; U.S. EPA, 2002). This approach is based on the expectation that the 6 pharmacodynamic aspect does not change between the various durations of within-day exposures 7 (<24 hr). Consistent with the Agency's policy (U.S. EPA, 2002), the AUC of a chemical for the 8 exposure scenario of the critical study should be determined initially using the PBPK model 9 (e.g., 6 hr/d, 5 d/wk); then the atmospheric concentration for continuous exposures (24 hr/d) 10 during lifetime or particular window of susceptibility yielding the same AUC or Cmax can be 11 determined by iterative simulation. 12 13 2.5.4. Dosimetric Adjustment Factor 14 In the RfC process, a dosimetric adjustment factor (DAF), intended to account for 15 pharmacokinetic differences between test species and humans, is applied to the duration-adjusted 16 POD (NOAEL, LOAEL, or BMC) to derive HECs (U.S. EPA, 1994). The DAF depends on the 17 nature of the inhaled toxicant and MOA as well as the endpoint (local effects vs. systemic 18 effects). The dosimetry data in the test animals and humans (e.g., deposition data, region- 19 specific dosimetry, blood concentration of systemic toxicants), if available, can help estimate the 20 DAF. In the absence of such data, knowledge of critical parameters or mathematical models in 21 the test species and humans can be useful in estimating the DAF. 22 For highly reactive or water-soluble gases that do not significantly accumulate in blood 23 (e.g., hydrogen fluoride, chlorine, formaldehyde, volatile organic acids, and esters), the DAF is 24 derived using simulated delivery of chemical to different regions of the respiratory tract, based 25 on regional mass transfer coefficients, as well as differences in surface area and ventilation rates 26 (U.S. EPA, 1994). For gases that are water soluble with some blood accumulation (e.g., acetone, 27 ethyl acetate, ozone, sulfur dioxide, propanol, isoamyl alcohol) and potential for both respiratory 28 and remote effects, the DAF could be calculated on the basis of uptake, defined by overall mass 29 transfer coefficient, as well as flow-limited perfusion distribution models. The DAF for poorly 30 water-soluble gases that cause remote effects (e.g., xylene, toluene, styrene) is determined on the 31 basis of blood:air partition coefficients and flow-limited perfusion models (U.S. EPA, 1994). 6/29/05 2-10 DRAFT—DO NOT CITE OR QUOTE ------- 1 A variety of computational tools are available for determining the uptake and deposition 2 of gases and particulates in nasal pathways and the respiratory tract (Kimbell et al., 1993; 3 Jarabek, 1994; Asgharian et al., 1995; Tran et al., 1997; Bush et al., 1998; Hanna et al., 2001, 4 Bogdanffy and Sarangapani, 2003; U.S. EPA, 2004). PBPK models have most frequently been 5 applied to systemically acting gases and vapors. They have also been applied, in conjunction 6 with other models (e.g., computational fluid dynamics), to more locally acting gases. Their use 7 (discussed above) in estimating dose metrics over time allows for a more refined duration 8 adjustment. Also, the other RfC approaches do not account for metabolism of the more reactive 9 gases, so PBPK modeling approaches would clearly be preferable for these compounds if 10 adequate data are available. There is little doubt that further applications of PBPK models to the 11 more reactive gases and agents will be developed and realized in the near future. 12 13 2.5.5. Uncertainty Factors: Role of PBPK Models 14 The uncertainty/variability factors used in RfC derivation include one or more of the 15 following: an interspecies uncertainty factor (IUF) of 3 to account for possible pharmacodynamic 16 differences; an intraspecies variability factor (IVF) of 10 for variability in kinetics and dynamics 17 among humans; subchronic-to-chronic extrapolation factor of 10; a LOAEL-to-NOAEL factor of 18 10; and a database deficiency factor of 1 to 10 (U. S. EPA, 1994; Jarabek, 1994). The total of all 19 uncertainty factors generally should not exceed 3,000 (U.S. EPA, 2002). With the exception of 20 the IVF, the magnitude of the factors listed above cannot be estimated using pharmacokinetic 21 data or PBPK models alone, although they can be determined either by conducting appropriate 22 experiments or by using integrated pharmacokinetic-pharmacodynamic models. Experiments to 23 estimate the IVF for each toxic chemical in each population of interest is neither plausible nor 24 ethical. It has been possible, however, to estimate the IVF by analyzing available data on 25 pharmaceutical products collected in human volunteers and patients over the years (Ham's et al., 26 1999; Silverman et al., 1999). 27 The IVF of 10 conventionally used in RfC derivation implies that for the same level of 28 response or nonresponse, the potential doses among individuals may differ by as much as—but 29 not more than—an order of magnitude. The factor of 10 is a composite of the interindividual 30 differences in pharmacokinetics and pharmacodynamics (Renwick and Lazarus, 1998). The 31 magnitude of the pharmacokinetic component of the IVF can be estimated with the use of PBPK 32 models and applied in conjunction with the remaining interindividual pharmacodynamic 6/29/05 2-11 DRAFT—DO NOT CITE OR QUOTE ------- 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 variability factor of 3, and this is entirely consistent with what is done under the RfC process to account for dosimetry and pharmacodynamic differences between species (Jarabek, 1994; U.S. EPA, 2002). The PBPK models are also potentially useful for estimating the magnitude of the pharmacokinetic component of the uncertainty factors for LOAEL-to-NOAEL and subchronic- to-chronic extrapolations, but the feasibility of such estimations has not yet been explored systematically (Thomas et al., 1996b). 2.5.6. Summary The potential role of the PBPK model in the RfC process is summarized in Box 2-2. Box 2-2. Role of PBPK models in the RfC process • Route-to-route extrapolation of the POD • Duration adjustment calculation • Dosimetric adjustment factor • Intraspecies variability factor (PK component) 2.6. ROLE OF PBPK MODELS IN RfD DERIVATION 2.6.1. Reference Dose An RfD is an estimate of a daily exposure to the human population (including sensitive subgroups) that is likely to be without an appreciable risk of deleterious effects during a lifetime (Barnes and Dourson 1988; Dourson et al., 1992). It is expressed in units of milligrams per kilogram per day. An RfD is calculated as follows: RfD = POD/UF where: POD - NOAEL, LOAEL, or BMD UF = uncertainty factors related to extrapolations associated with the POD (i.e., interspecies extrapolation, intraspecies extrapolation, subchronic-to-chronic extrapolation, LOAEL-to-NOAEL extrapolation) or incompleteness of the database. An RfD derivation begins with the identification of the POD for the critical effect. Subsequently, the uncertainty factors are applied as appropriate. Regarding the RfD process, PBPK models are potentially useful in determining the POD as well as the extrapolation factors, as described in Sections 2.6.2 and 2.6.3. 6/29/05 2-12 DRAFT—DO NOT CITE OR QUOTE ------- 1 2.6.2. Point of Departure 2 Where possible, the RfD derivation uses an oral NOAEL, LOAEL, or BMD as the POD. 3 The oral route-specific NOAEL and LOAEL correspond to experimentally tested doses, whereas 4 the BMD is obtained from statistical modeling of dose-response data (U.S. EPA, 2000a). When 5 the LOAEL or NOAEL has not been identified in an oral dosing study but has been obtained for 6 another route of exposure, the route-to-route dose conversion in the past assumed that absorption 7 by both routes is 100% and that equal potential doses administered via inhalation and oral route 8 are equitoxic (Pauluhn, 2003). In other words, this overly simplistic approach assumed that the 9 rates of absorption, distribution, metabolism, excretion, and tissue dose of toxic moiety of 10 chemicals are identical, regardless of the exposure route, particularly if the same dose (mg/kg/d) 11 is administered by both routes. The data on route-specific fraction absorbed, when available, 12 have been used to improve the scientific basis of the route-to-route extrapolation process (e.g., 13 U.S. EPA, 1999a). 14 When the oral NOAEL, LOAEL, or BMD is unavailable for a given route of exposure, 15 PBPK models can be particularly useful in deriving such values on the basis of results obtained 16 for other dosing routes (e.g., inhalation, intravenous, dermal). PBPK models facilitate the 17 conduct of route-to-route extrapolation (e.g., inhalation to oral) to establish the oral route- 18 specific POD (NOAEL, LOAEL, BMD), based on equivalent internal dose. The use of PBPK 19 models represents a significant improvement in the scientific basis of the conventional route-to- 20 route extrapolation approach, because it takes into account the route-specific rates and magnitude 21 of absorption, clearance, and first-pass effects. 22 PBPK models, however, are not able to establish an oral POD in the absence of some 23 route-specific experimental dose-response and kinetic data. Development of a biologically based 24 dose-response (BBDR) model linked with a PBPK model could potentially improve the 25 quantification of the dose-response and POD for RfD derivation. There has been some success 26 in developing BBDR models for simple adverse effects (e.g., cholinesterase inhibition, 27 cytotoxicity, hematoxicity) (Gearhart et al., 1990, 1994; Reitz et al., 1990a; Cox, 1996; Ashani 28 and Pistinner, 2004), but these models are not routinely used to estimate PODs for RfD 29 derivation because of the lack of data needed to calibrate and test the models. New technologies 30 (e.g., toxicogenomics) offer the potential for generating the needed data, and more integrated 31 PBPK and BBDR models may be an attainable goal in the near future. 32 6/29/05 2-13 DRAFT—DO NOT CITE OR QUOTE ------- 1 2.6.3. Uncertainty Factors 2 The uncertainty factors used in RfD derivation include extrapolations across species 3 (interspecies) and across duration of exposure (subchronic to chronic), degree of response (e.g., 4 LOAEL-to-NOAEL), variability within a population (intraspecies variability), and to account for 5 poor or absent data in the database (database deficiency). If the NOAEL for a chemical with an 6 adequate database has been identified in a chronic study, only the IUF and IVF are used in the 7 assessment. The IUF of 10 conventionally used in RfD derivation addresses the possibility that 8 the same level of response or nonresponse is associated with a HED that is 10 times lower than 9 the dose administered to animals. Similarly, the FVF of 10 is presumed to account for both the 10 differences in tissue dosimetry and tissue response within a human population. 11 If in vivo pharmacokinetic and pharmacodynamic data are available for humans and the 12 test species as well as for the various subgroups of the population, chemical-specific adjustment 13 factors can be derived to replace the default uncertainty factors. The IUF and IVF may also be 14 estimated from an understanding of the key pharmacokinetic and pharmacodynamic 15 determinants of a chemical's disposition in the body (e.g., isoenzyme levels, enzyme activity 16 levels, tissue volumes, breathing rates, cell proliferation rates) (e.g., Dome et al., 2001a, b, 2002; 17 Walton et al., 2001). Extrapolating among species or estimating interindividual variability must 18 be performed carefully because the magnitude of variability associated with individual 19 determinants is neither cumulative nor additive. A one-by-one evaluation of the determinants of 20 interspecies differences or interindividual variability is useful for understanding the mechanistic 21 basis and their relative importance, but such an isolated approach can easily lead to unrealistic 22 estimates of the overall magnitude of an FVF or IUF. The net impact of various determinants on 23 the IVF and IUF is more properly evaluated by integrating the available information with a 24 PBPK or BBDR model. 25 PBPK models are useful for characterizing the magnitude of the pharmacokinetic 26 component of the IUF as well as the IVF used in the RfD process. When using PBPK models to 27 adjust for pharmacokinetic differences, a factor of 3 is generally used to account for the potential 28 interspecies differences and interindividual variability in pharmacodynamics, as with the RfC 29 process (Jarabek et al., 1995; Clewell et al., 2002; U.S. EPA, 2003). However, chemical-specific 30 information on the pharmacodynamic aspect of inter- and intraspecies differences may inform on 31 further reduction of these uncertainty factors. It should be recognized that in their current state 6/29/05 2-14 DRAFT—DO NOT CITE OR QUOTE ------- 1 of advancement, PBPK and BBDR models are not suitable for characterizing the magnitude of 2 LOAEL-NOAEL and subchronic-chronic factors. 3 Although a thoroughly characterized BBDR model might address the LOAEL-NOAEL 4 issue, it is unlikely that such models will be capable of extrapolating from subchronic to chronic 5 effects because of the multiplicity of feedback systems and plasticity within organisms, nor will 6 such models be capable of accounting for the potential for different adverse effects over the 7 lifespan. 8 9 2.6.4. Summary 10 PBPK models, by facilitating the Box 2-3. Role of PBPK models in the RfD process 11 simulation of tissue dose of the toxic Route-to-route extrapolation Interspecies uncertainty factor (PK component) Intraspecies variability factor (PK component) 12 moiety of chemicals, address specific areas 13 of uncertainty associated with derivation of 14 the RfD, as shown in Box 2-3. 15 16 2.7. ROLE OF PBPK MODELS IN CANCER RISK ASSESSMENT 17 The dose-response assessment portion of cancer risk assessment may vary, depending on 18 MO A considerations. It can involve the determination of a slope factor based on linear 19 extrapolation from the POD (i.e., high-dose to low-dose extrapolation) or it may involve the use 20 of the POD for nonlinear analysis (U.S. EPA, 2005a). Either approach may also require 21 extrapolation of animal data to humans, hi addition, some of the assessments may require route- 22 to-route extrapolations. The role of PBPK models in conducting these extrapolations is 23 discussed in the following sections (2.7.1 through 2.7.4). 24 25 2.7.1. High-Dose to Low-Dose Extrapolation 26 The oral slope factor or IUR has frequently been determined by modeling the relationship 27 between the cancer response and the administered dose (or exposure concentration) (U.S. EPA, 28 2005a). According to the revised cancer guidelines, however, either a nonlinear (i.e., RfC or 29 RfD) or linear (i.e., unit risk estimate) extrapolation based on the POD can be conducted, as 30 justified by the MOA of the carcinogen (U.S. EPA, 2005a). Further, the use of internal dose or 31 delivered dose in such analysis has been encouraged. 6/29/05 2-15 DRAFT—DO NOT CITE OR QUOTE ------- 1 Because high doses of chemicals are often administered in rodent cancer bioassays, the 2 tumors observed in such studies are not always directly proportional to the exposure dose. 3 Rather, the cancer dose-response relationships appear complex, in part due to nonlinearity in the 4 pharmacokinetic processes occurring at high doses. In other words, the target tissue dose of the 5 toxic moiety is often disproportional to the administered doses used in animal bioassays (Figure 6 2-1). Therefore, dose-response analysis based on the appropriate dose metric often results in the 7 linearization of the relationship (Andersen et al., 1987; Clewell et al., 1995, 2002). The slope 8 factor derived using the dose metric versus the response curve will have units of (dose metric)"1; 9 similarly, NOAELs are converted using PBPK models to the dose metric at which no significant 10 incidence of cancer is expected on the basis of MO A of the chemical and dose-response data. 11 Linkage of PBPK models with BBDR models (e.g., clonal expansion and progression 12 models as well as two-stage models) would represent the ideal framework for characterizing the 13 health risk associated with human exposure to chemical carcinogens. However, such models still 14 continue to be developed and improved upon, and currently there is not a standard model that is 15 used or recommended (U.S. EPA, 2005a). 16 17 2.7.2. Interspecies Extrapolation 18 The default procedure for conducting interspecies extrapolation (e.g., rat to humans) of 19 the oral dose of carcinogens involves scaling in proportion to body weight to the 0.75 power 20 (U.S. EPA, 2005a). This scaling approach is based on the observation that several physiological, 21 anatomical, and biochemical processes in mammalian species are closely related to body surface 22 area (Voisin et al., 1990; U.S. EPA, 2002). In other words, this default approach implies that the 23 delivered doses (following oral exposure) are related to applied dose by 0.75 power of body 24 weight, independent of the species. For gases and particulates, however, the default procedure 25 for interspecies extrapolation would involve the derivation of a HEC, as described in Section 26 2.4.4 (U.S. EPA, 1994; Jarabek, 1995a, b). 27 The nature and slope of the dose-response relationship for carcinogens may not be 28 identical in test species and humans due to pharmacokinetic and pharmacodynamic differences 29 (Monro, 1994). Interspecies extrapolations of an equivalent dose of carcinogens can be 30 conducted if appropriate data (tissue or blood concentrations of potential toxic moiety or 31 promutagenic adducts) are available in both the test species and humans. In the absence of such 32 data, PBPK models can be used to characterize the relationship between applied dose and 6/29/05 2-16 DRAFT—DO NOT CITE OR QUOTE ------- 1 delivered dose of carcinogens in the species of interest for subsequent extrapolation to humans 2 (Andersen et al., 1987). 3 4 2.7.3. Route-to-Route Extrapolation 5 Route-to-route extrapolation is necessary if the cancer dose-response data were obtained 6 in test species exposed by a route different from the anticipated human exposure and the effect is 7 considered pertinent to the human exposure route. Route-to-route extrapolation is considered 8 appropriate for a chemical that induces tumors at a site different from the portal of entry and if 9 the chemical is absorbed to give an internal dose. Simplistic route-to-route extrapolations are 10 performed by assuming that the relationship between applied dose and tissue dose of toxic 11 moiety (dose metric) is the same, regardless of the exposure route. The uncertainty associated 12 with this approach is that the first-pass effect as well as the rates and extent of absorption and 13 metabolism may vary from one route to another (Pauluhn, 2003). The data on route-specific 14 fraction absorbed are now used to improve the scientific basis of the route-to-route extrapolation 15 process (e.g., U.S. EPA, 1999a). 16 PBPK models, by accounting for the route-specific rate and magnitude of absorption, 17 first-pass effect, and metabolism, facilitate the conduct of route-to-route extrapolation (Clewell 18 and Andersen, 1994). The slope factor or the NOAEL associated with an exposure route can be 19 translated into applied dose by another route on the basis of equal delivered doses. Basically, by 20 simulating the tissue dose of toxic moiety associated with the applied doses given by different 21 routes, PBPK models facilitate conducting route-to-route extrapolation of the NOAEL or slope 22 factor of chemical carcinogens (Gerrity et al., 1990; U.S. EPA, 2000b). 23 24 2.7.4. Intraspecies Variability 25 Intraspecies variability in pharmacokinetics or pharmacodynamics has not usually been 26 considered in the context of cancer risk assessment. In characterizing risk, the CSF has been 27 used without adjustment for susceptible populations. The newer supplemental guidance suggests 28 the use of an adjustment factor to the cancer slope or unit risk value to account for enhanced 29 susceptibility related to early -life exposures of neonates and young children, particularly for 30 carcinogens exhibiting a mutagenic MOA (U.S. EPA, 2005b). Further, when assessing the less- 31 than-lifetime exposures occurring in childhood, the guidelines stipulate adjustments for adult- 6/29/05 2-17 DRAFT—DO NOT CITE OR QUOTE ------- 1 children differences in exposure factors (e.g., skin surface area, drinking water ingestion rate) 2 (U.S. EPA, 2005b). 3 PBPK models are potentially of use in evaluating the pharmacokinetic basis of the adult- 4 children differences in tissue dose of carcinogens (Price et al., 2003b; Gentry et al., 2003; 5 Ginsberg et al., 2004). However, the quantitation of the tissue dose differences between adults 6 and children would not be sufficient to account for pharmacodynamic differences related to 7 early-life exposures of neonates and children. PBPK models can still be of use in determining 8 the adjustment factor for early-life exposures of neonates and young children on the basis of 9 equivalent internal doses. 10 11 2.7.5. Summary 12 PBPK models, by facilitating the 10 .,. _. , f. • • 4, f * Route-to-route extrapolation 13 simulation of tissue dose of toxic moiety of T . * . „ , ,- • „ •» Interspecies extrapolations of pharmacokinetically 14 chemical carcinogens, play an important 15 role in reducing the uncertainties associated Box 2-4. Role of PBPK models in cancer risk assessment • High-dose to low-dose extrapolation equivalent doses 16 with some of the extrapolations used in the cancer risk assessment process (Box 2-4). 17 18 2.8. USE OF PHARMACOKINETIC DATA AND MODELS IN EXPOSURE 19 ASSESSMENT 20 2.8.1. Conventional Approaches 21 The conventional approach to exposure assessment involves the calculation of potential 22 dose for each route of exposure, with knowledge of concentration of the chemical in the medium, 23 frequency and duration of exposure, rate of contact with the medium, and body weight of the 24 individual (Paustenbach, 2000). With increased data availability, however, absorbed dose can be 25 calculated (U.S. EPA, 1992). For calculating absorbed dose, pharmacokinetic data such as time- 26 course data on concentration or total quantity in alveolar air, urine, or blood would be required 27 (Paustenbach, 2000). The conversion of biomarker data, absorbed dose, or potential dose into 28 delivered dose is not straightforward and requires the use of appropriate modeling techniques. 29 The need for the use of pharmacokinetic models is further emphasized by the fact that the tissue 30 dose is not always directly proportional to the exposure concentration or potential dose. 31 6/29/05 2-18 DRAFT—DO NOT CITE OR QUOTE ------- 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 2.8.2. Role of PBPK Models PBPK models allow the conversion of potential dose or exposure concentration to tissue dose, which can then be used for risk characterization purposes. As the dose-response relationships are being increasingly established on the basis of tissue dose, it is also essential that the potential doses calculated during exposure assessments be translated into tissue doses to facilitate risk characterization. PBPK models are additionally useful in investigating and establishing the relationship between biomonitoring or biomarker data and potential dose as well as the relationship between biomarker levels and tissue dose of chemicals by iterative simulation (e.g., Fennell et al., 1992; Krishnan et al., 1992; Csanady et al., 1996; Timchalk et al., 2001). This ability of PBPK models has been explored for establishing biological exposure indices (e.g., breath, blood, or breath concentrations) associated with worker exposure to threshold limit values of solvents (Perbellini et al., 1990; Leung, 1992; Kumagai and Matsunaga, 1995; Thomas et al., 1996b; Droz et al., 1999). PBPK models also offer a framework for reconstructing human exposure over a period of time (Vinegar et al., 1990; Roy and Georgopoulos, 1998; Canuel et al., 2000). Further, such dose reconstructions and tissue dose resulting from multiroute (oral, inhalation, dermal) and multimedia (air, water, food, soil) exposures can be computed with the use of PBPK models (e.g., Georgopoulos et al., 1994; Roy et al., 1996; Rao and Ginsberg, 1997; Corley et al., 2000; Levesque et al., 2002). The net tissue dose associated with a particular exposure situation (multiroute and multisource) can then be obtained with PBPK models and used for risk characterization purposes. 2.8.3. Summary The potential roles of PBPK models in exposure assessment are summarized in Box 2-5. Box 2-5. Role of PBPK models in exposure assessment • Conversion of potential dose into tissue dose • Tissue dose estimation for aggregate exposures • Dose reconstruction based on exposure biomarkers 2.9. PHARMACOKINETIC DATA NEEDS IN RISK ASSESSMENT: SUMMARY Adverse tissue responses are more directly and closely related to tissue dose of the toxic moiety than is the concentration of the parent chemical in the environment. Therefore, there is no question that the scientific basis of, and confidence in, risk assessments are enhanced by the use of tissue dose information. Because the tissue dose is not directly and simply related to exposure concentration or potential dose at all tunes, there is some uncertainty regarding the use 6/29/05 2-19 DRAFT—DO NOT CITE OR QUOTE ------- 1 of potential dose in risk assessment. If the relationship between potential dose and tissue dose of 2 the toxic moiety of chemicals is linear and proportional at all dose levels used in animals and 3 humans, then there is no need for PBPK models. However, this is not usually the case. 4 Therefore, the PBPK models are used for addressing specific areas of uncertainty in dose- 5 response assessment and exposure assessment on the basis of the simulation of tissue dose of 6 potential toxic moiety. 7 8 In the context of dose-response assessment, PBPK models are needed for conducting 9 10 1. Route-to-route extrapolation of the POD (RfC, RfD, and unit risk estimates), 11 2. Duration adjustment of the POD (RfC derivation), 12 3. High-dose to low-dose extrapolation (unit risk estimates), 13 4. Interspecies extrapolation of pharmacokinetically equivalent doses (RfD, RfC, and unit 14 risk estimates), and 15 5. Estimation of the pharmacokinetic component of the IVF (RfC and RfD derivation). 16 17 In the context of exposure assessment, PBPK models are needed for 18 19 1. Converting potential dose into tissue dose, 20 2. Calculating tissue dose associated with multiroute and multimedia exposures, and 21 3. Relating biomarker data to tissue dose and potential dose by exposure reconstruction. 22 23 The PBPK models intended for use in dose-response and exposure assessments should be 24 evaluated for their adequacy. The next chapter presents the process and criteria for evaluating 25 PBPK models intended for use in risk assessment. 6/29/05 2-20 DRAFT—DO NOT CITE OR QUOTE ------- 1 3. EVALUATION OF PBPK MODELS INTENDED FOR USE IN 2 RISK ASSESSMENT 3 4 PBPK models intended for risk assessment applications should be evaluated for their 5 adequacy. There are no published documents that specify either model validation standards or 6 criteria of adequate PBPK models. However, several publications have dealt with good 7 modeling practices as well as approaches for evaluating and documenting biological models 8 intended for risk or safety assessments (Yates, 1978; Andersen et al., 1995; Clark et al., 2004). 9 Evaluation of PBPK models intended for risk assessment applications should be undertaken 10 along the lines of model purpose, model structure, mathematical representation, parameter 11 estimation (calibration), and computer implementation aspects, as detailed in the following 12 sections. 13 14 3.1. MODEL PURPOSE 15 The purpose for which a PBPK model is developed basically determines its structure and 16 level of detail. Not all PBPK models are developed for the purpose of supporting risk 17 assessments. Some are actually used as a tool for hypothesis testing for particular modes of 18 action and for guiding improved experimental design. The structure of a PBPK model used 19 strictly for responding to the needs of a curiosity-driven researcher may not serve the intent and 20 purposes of a risk assessor. PBPK models, therefore, are of varying complexities and 21 capabilities, according to the intended purpose. PBPK models, in general, are developed for one 22 of the following three purposes: 23 1. Integrate diverse sets of pharmacokinetic data on a particular chemical, 24 2. Investigate the pharmacokinetic basis of toxicity of a chemical that appears complex 25 at the administered dose level, and 26 3. Predict tissue dosimetry for situations other than what has been or could be tested 27 experimentally (see Section 3.6). 28 29 Models that permit the integration of diverse sets of data are useful in uncovering 30 mechanistic determinants of pharmacokinetics of specific chemicals (e.g., Clewell and Andersen, 31 1997; Haddad et al., 1998). However, PBPK models resulting from such efforts are not routinely 32 evaluated for their predictive value but still are useful for the purpose of uncovering unusual 6/29/05 3-1 DRAFT—DO NOT CITE OR QUOTE ------- 1 determinants of pharmacokinetics (e.g., suicide inhibition, diffusion-limited uptake). On the 2 other hand, PBPK models developed to improve the understanding of toxicological behavior of 3 chemicals that appears complex at the administered dose level (e.g., species, sex, or route 4 differences in response) may be either predictive or integrative. Typically, the PBPK models of 5 use to the risk assessor are those that have predictive value at the quantitative level with respect 6 to the production or pharmacokinetics of potential toxic moieties (parent chemical or 7 metabolite). 8 9 A PBPK model capable of predicting the pharmacokinetics and tissue dose of the 10 potential toxic moiety of a chemical is the preferred one for risk assessment applications. 11 12 3.2. MODEL STRUCTURE 13 The structure of a PBPK model in large part depends upon the purpose for which the 14 model is developed and the philosophy of the modeler. There is virtually no limit to the number 15 and size of compartments in a model intended to describe molecular/cellular events. Parsimony 16 in selecting model structures, however, is an important and guiding principle in developing 17 models for use in risk assessments. The complexity of PBPK models used in risk assessment is 18 often constrained by limited data available to calibrate and test the model, and the need for risk 19 assessors to defend the model assumptions and the values derived from model simulations. 20 The simplest conceptual model represents the organism as a one-compartment system. 21 PBPK models differ from single-compartment models by representing many more physiological 22 and biochemical processes that are relevant to the toxicokinetics of the chemical in question. 23 This complexity is represented by differentiating or lumping tissues into specific compartments 24 each with a unique set of physiological (i.e., blood flows) and biochemical parameters (i.e., 25 partition coefficients). A PBPK model intended for risk assessment applications should, 26 preferably, include the target organ as one of the compartments. Minimally, it should facilitate 27 the calculation of blood concentration, which is often used as a surrogate for tissue 28 concentrations. Major portals of entry (e.g., lung, gastrointestinal tract), storage organs (e.g., 29 adipose tissue), metabolism/transformation sites (e.g., liver, kidney) as well as elimination routes 30 (e.g., renal, pulmonary, fecal) should be included. It is often acceptable to mathematically 31 describe absorption, distribution, metabolism and excretion (ADME) of chemicals in PBPK 32 models without physically representing the tissues where these processes occur (Krishnan and 6/29/05 3-2 DRAFT—DO NOT CITE OR QUOTE ------- 1 Andersen 2001), provided that this lack of physical representation does not interfere with a 2 model's use as an extrapolation tool. In some cases, a compartment is subdivided into several 3 sub-compartments based on mechanistic and biological considerations. For example, the liver 4 can be divided into separate compartments depending on the localization of enzymatic activity. 5 Figure 3-1 represents examples of PBPK model structures that have been commonly used to 6 simulate the kinetics of volatile and nonvolatile substances. Note that all these models facilitate 7 the simulation of the concentration of chemicals or their metabolites in the target organ or a 8 surrogate tissue (usually blood). 9 It is important to realize that compartments in PBPK models are usually assumed to be 10 homogenously and completely mixed reactors. This means that the concentration of the chemical 11 anywhere in the tissue is the same and equal to the concentration of the chemical as it leaves the 12 tissue in venous blood. This assumption is necessary to simplify the differential equations 13 representing the mass balance of the chemical in the tissue. The numerical solution of the final 14 set of equations representing the PBPK model is further simplified because of the complete- 15 mixing assumption. 16 The structure of a PBPK model intended for risk assessment applications should contain 17 the target organ (or a surrogate tissue) as well as compartments representing tissues of unique 18 physiological and biochemical relevance to the pharmacokinetics of the chemical in question. 19 20 3.3. MATHEMATICAL REPRESENTATION 21 In PBPK modeling, each tissue compartment is generally described with a mass balance 22 differential equation (MBDE) that consists of a series of clearance terms with units of volume 23 per time, i.e., liters per hour or milliliters per minute. The clearance terms, in most cases, relate 24 to tissue uptake, tissue-to-blood transfer, metabolism, or excretion of chemicals. The uptake of a 25 chemical in systemic circulation by a tissue is described according to Pick's law of simple 26 diffusion, which states that the flux of a chemical is proportional to its concentration gradient. 27 Descriptions of passive and blood flow-limited uptake have been used successfully in many of 28 the past efforts in PBPK modeling that dealt with small-molecular-weight organics. For high- 29 molecular-weight compounds, however, membrane diffusion is often the rate-limiting process, 30 and in such cases chemical uptake is described with differential equations for the tissue blood 31 and cellular matrix subcompartments (Rowland, 1985; Leung, 1991; Andersen, 1995; Krishnan 32 and Andersen, 2001) (Tables 3-1 and 3-2). 6/29/05 3-3 DRAFT—DO NOT CITE OR QUOTE ------- B Inhaled air Expired air Inhaled air Expired air 2 3 4 5 6 7 8 9 10 ^ r ' A LUNG RICHLY TISSUES POORLY TISSUES 1 I\/CD fc * ^ ' Motahnlici LUNGS 0u tichangt •-Jfc LUNGS IMOIXXC D Metabolism Figure 3-1. Conceptual representations of PBPK models for (A) toluene, (B) ethylene oxide, (C) 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD), and (D) vinyl acetate. The input and output arrows associated with individual compartments represent arterial and venous blood flows. Note that model D contains 5 compartments for the nasal cavity alone (the site of toxicity). From Bogdanffy et al. (1999), Krishnan et al. (1992), and Andersen et al. (1993). 6/29/05 3-4 DRAFT—DO NOT CITE OR QUOTE ------- 1 2 Table 3-1. Equations of a four-compartment PBPK model to simulate the inhalation exposure of volatile organic compounds Tissue compartments Equations8 Arterial blood Ca ^ Pi, Liver dt l - '"-1' Km + Cv. . * n— i dA, Fat dA, dt Af = J n Cv, =• dA, Richly perfused tissues dt r - ^ ~ y Cv = ^-V = r" Poorly perfused tissues dASn ~dt A, = sn -Cv M Sft- dA, 2* ?-> _ Sn ~ 6/29/05 3-5 DRAFT—DO NOT CITE OR QUOTE ------- 1 2 3 4 5 6 7 8 9 10 11 12 13 14 Table 3-1. Equations of a four-compartment PBPK model to simulate the inhalation exposure of volatile organic compounds (continued) Tissue Compartments Venous blood Alveolar air Equations8 ^ QlxCvln+QfxCvfn+QrxCvrn+QSxCv,n " Qc Calv - Ca" " P» 1 Equations are from Ramsey and Andersen (1984). bThe arterial blood equation in this example is used for chemicals that reach rapid equilibrium in blood, such as highly fat-soluble volatile chemicals. In other cases, a detailed mass-balance equation for the arterial blood may be needed. A = amount (mg) a = arterial blood alv = alveolar air b = bloodiair C = concentration (mg/L or mmol/L) c = cardiac output f=fat inh = inhaled air Km = Michaelis Menten affinity constant (mg/L) 1 = liver n = current time n-1 = previous simulation time P = partition coefficient p = pulmonary ventilation Q = How rate (L/hr'1) r = richly perfused tissues s = slowly perfused tissues t = tissue:blood V = Volume (L) v = mixed venous blood vf = venous blood leaving fat vl = venous blood leaving liver Vraax = maximal velocity of enzymatic reaction (mg/hr'1) vr = venous blood leaving richly perfused tissue vs = venous blood leaving poorly perfused tissue Table 3-2. Equations used for describing diffusion-limited uptake in PBPK models Subcompartments Tissue blood Cellular matrix Equations f\ ' ~ JZrt ^ \ m ~~ nut ) ~ I-* VT- 1 ^ \1 "7 / M/ ^^• = Mx(c,-c2) A = amount (mg) C = concentration (mg/L or mmol/L) in = inflow out = outflow PA = permeability-area coefficient Q = flow rate (L/hr'1) tl = tissue blood t2 = cellular matrix V = volume (L) 6/29/05 3-6 DRAFT—DO NOT CITE OR QUOTE ------- 1 The rate of the amount of chemical consumed by macromolecular binding process is 2 described in PBPK models as a second-order reaction or by using equations based on reversible 3 equilibrium relationship. The rate of metabolism in PBPK models has been described as a first- 4 order, second-order, or saturable process. Conjugation reactions, on the other hand, are 5 traditionally described as a second-order process with respect to the concentrations of the 6 cofactor and the chemical. Alternatively, descriptions based on ping-pong mechanism have also 7 been used successfully. In each case, the reason for using a particular description should be 8 clearly provided. PBPK models using particular mathematical descriptions of tissue uptake, 9 metabolism, and binding without any justification cannot be used confidently for risk assessment 10 applications. For example, if enzyme-mediated metabolism is described as a first-order process 11 in a PBPK model, the scientific rationale for employing such a description is needed before the 12 model can be used for purposes of extrapolation and prediction. Because PBPK models are 13 simplified representations of the real systems, the actual details and complexity of the 14 physiological and biochemical processes are not reflected by the equations used. Depending on 15 the level of detail required and the objective of the modeling effort, appropriate descriptions of 16 the biochemical processes can be included in these models. 17 - The MBDEs in PBPK models should have units of mass per time (e.g., mg/hr) or 18 concentration per time (mg/L-hr'1), and the type of equations chosen to describe tissue uptake, 19 binding, and metabolism should be justified. 20 21 3.4. PARAMETER ESTIMATION 22 Knowledge of chemical-specific and species-specific parameters would be required to 23 solve the equations constituting PBPK models. For a model to be used for estimating 24 interindividual differences in tissue dosimetry, knowledge of the distributions of input 25 parameters is essential. However, for all other purposes, the knowledge of the average value or 26 the range of plausible values of model parameters is sufficient. Typically, PBPK models require 27 the numerical values of physiological parameters, such as alveolar ventilation rate, cardiac 28 output, tissue blood flow rates and tissue volumes, and clearance parameters related to renal, 29 hepatic, and biliary excretion pathways, as well as certain partition coefficients (blood:air, 30 skin:water, skin:air, and tissue:blood). Additional parameters (e.g., tissue DNA levels, 31 hematocrit, number and concentration of binding proteins) may be required in some cases. 32 6/29/05 3-7 DRAFT—DO NOT CITE OR QUOTE ------- 1 3.4.1. Physiological Parameters 2 The physiological parameters used in PBPK models should either correspond to those 3 obtained in the experimental pharmacokinetic study or be within the range of plausible values for 4 the species and life stage. Even though peer-reviewed compilations of ranges and reference 5 values of physiological parameters for adult animals and humans are available (Arms and Travis, 6 1988; Leggett and Williams, 1991; Fiserova-Bergerova, 1995; Brown et al., 1997; Davies and 7 Morris, 1997) (Tables 3-3 through 3-6), this is not the case with respect to physiological 8 parameters for specific subgroups of populations (e.g., developing and lactating animals, 9 pregnant women, children). There are some reports on physiological parameters for specific 10 subgroups but no compilations as yet of definitive ranges or reference values (Luecke et al., 11 1994; Schoeffner et al., 1999; Haddad et al., 2001a; Hattis et al., 2003; Price et al., 2003a, b; 12 Pelekisetal., 2003). 13 In PBPK models for organic chemicals, the sum total of the volumes of compartments 14 corresponding to soft tissues should be smaller than the body weight, usually about 91% of the 15 body weight (100% [body weight] - 9% [weight of skeletal/structural components as percent 16 body weight]). Even though the tissue volumes (L) are needed for PBPK modeling, tissue 17 weights (kg) are usually used with the assumption of unit density (L = Kg). This assumption, 18 which may seem questionable, is inconsequential for practical reasons, particularly with respect 19 to the application of PBPK models in risk assessment. 20 The tissue flow rates in the model should add up to cardiac output. Fundamentally, 21 maintaining the mass balance in PBPK models requires that the sum of the flows to the 22 compartments be equal to the cardiac output. The ratio of cardiac output to alveolar ventilation 23 rate is around 1 in a resting individual (e.g., Andersen et al., 1987). The specification of cardiac 24 output independent of the value of ventilation rate is unacceptable, particularly if the their ratio 25 (ventilation:perfusion ratio) is not in the normal physiological range. Frequently in PBPK 26 models, ventilation rate, cardiac output, and tissue perfusion rates and tissue volumes are 27 specified for an individual animal or human being simulated. For predicting kinetics in another 28 animal or human with a different body weight, the physiological parameters are calculated anew. 29 To simplify this process, all tissue volumes are expressed as fractions of body weight such that 30 for any given body weight, the volumes in liters can be readily calculated by multiplying the 31 body weight by the corresponding fractional value. Similarly, because the cardiac output and 32 alveolar ventilation rate are related to body surface rather than body weight, the PBPK models 33 6/29/05 3-8 DRAFT—DO NOT CITE OR QUOTE ------- Table 3-3. Reference physiological parameters for mice, rats, and humans Physiological parameters Body weight (BW) (kg) Tissue volume (fraction of BW) Liver Fat Organs Muscle and skin Cardiac output (Qc) (L/min) Tissue perfusion (fraction of Qc) Liver Fat Organs Muscle and skin Minute volume (L/min) Alveolar ventilation (L/min) Mouse 0.025 0.055 0.1 0.05 0.7 0.017 0.25 0.09 0.51 0.15 0.037 0.025 Rat 0.25 0.04 0.07 0.05 0.75 0.083 0.25 0.09 0.51 0.15 0.174 0.117 Humans 70 0.026 0.19 0.05 0.62 6.2 0.26 0.05 0.44 0.25 7.5 5 2 3 4 Source: Travis and Hattemer-Frey (1991). 6 7 8 9 10 11 Table 3-4. Range of values of the volume and perfusion of select tissues in the mouse Tissue Adipose Brain Heart Kidneys Liver Lungs Muscle Skin Volume (% body weight) 5-1 4a 1.35-2.03 0.4-0.6 1.35-1.88 4.19-7.98 0.66-0.86 35.8-39.9 15.9-20.8 Regional blood flow (% cardiac output) 3.1-3.5 5.9-7.2 7-11.1 12.2-19.6 3.3-8.3 * Varies proportionately with body weight. Source: Brown et al. (1997). 6/29/05 3-9 DRAFT—DO NOT CITE OR QUOTE ------- 1 2 Table 3-5. Range of values of the volume and perfusion of select tissues in the rat Tissue Adipose Brain Heart Kidneys Liver Lungs Muscle Skin Volume (% body weight) 4.6-1 2a 0.38-0.83 0.27-0.4 0.49-0.91 2.14-5.16 0.37-0.61 . 35.4-45.5 15.8-23.6 Regional blood flow (% cardiac output) 1.5-2.6 4.5-5.1 9.5-19 13.1-22.1 11.1-17.8 3 4 5 6 a Varies proportionately with body weight. Source: Brown et al. (1997). Table 3-6. Range of values of perfusion of select tissues in humans 9 10 11 12 13 14 15 16 17 Tissue Adipose Brain Heart Kidneys Liver Muscle Skin Regional blood flow (% cardiac output) 3.7-11.8" 8.6-20.4 3.8-8 12.2-22.9 11-34.2 5.7^t2.2 3.3-8.6 a Varies proportionately with body weight. Source: Brown et al. (1997). specify these parameters as a power function of body weight, with the exponent varying from 0.67 to 0.75 (e.g., Andersen et al., 1987; Tardif et al., 1997). An acceptable PBPK model should contain tissue volumes, flow rates, and ventilation:perfusion ratios that are within physiological limits. Particularly, the sum total of 6/29/05 3-10 DRAFT—DO NOT CITE OR QUOTE ------- 1 the tissue volumes should not exceed the body weight, and the sum total of tissue blood flow rates 2 should equal cardiac output. 3 4 3.4.2. Partition Coefficients 5 The calibration of PBPK models for partition coefficients has sometimes been done using 6 in vivo data. In such cases, pharmacokinetic data collected following a single bolus dose or 7 repeated doses leading to steady state are analyzed to determine the tissue:blood partition 8 coefficients (Chen and Gross, 1979; Lam et al., 1982; Gabrielsson and Bondesson, 1987; Gallo 9 et al., 1987). The steady-state data provide the most straight forward data for model calibration; 10 however, they are valid only for tissues in which there are no significant binding or metabolic 11 processes. In case of significant level of metabolism or binding, the calculation of tissue:blood 12 partition coefficients should account for the amount of chemical consumed by such processes 13 (Chen and Gross, 1979). The estimation of partition coefficients for PBPK models from in vivo 14 data is acceptable as long as the same data set is not used for external evaluation later on. 15 The tissue:air, skin:water, skin:air, and blood:air partition coefficients required for PBPK 16 modeling of volatile organic chemicals are conveniently determined in vitro using vial 17 equilibration method (Sato and Nakajima, 1979; Gargas et al., 1989; Johanson and Dynesius, 18 1988; Fiserova-Bergerova and Diaz, 1986; Kaneko et al., 1994; Perbellini et al., 1995; Beliveau 19 and Krishnan, 2000). Tissue:blood partition coefficients for nonvolatile chemicals can be 20 determined in vitro using radioactive chemicals in ultrafiltration, equilibrium dialysis, or vial 21 equilibration procedure (Lin et al., 1982; Igari et al., 1983; Sultatos et al., 1990; Jepson et al., 22 1994; Murphy et al., 1995). The partition coefficients estimated by these in vitro methods are 23 acceptable, provided equilibrium is attained during the experimental conditions. The 24 experimenter/modeler should have conducted a time-course analysis to choose an appropriate 25 time point (at which equilibrium is attained) for determining partition coefficients in vitro. 26 Algorithms based on the consideration of solubility and binding of chemicals in 27 biological matrices have been developed and applied for predicting tissue:blood, tissue:air, and 28 blood: air partition coefficients of volatile organic chemicals. This approach requires knowledge 29 of tissue and blood composition in terms of lipid and water contents, and octanol: water or 30 oil:water partition coefficients of the chemical as well as the binding association constants, if 31 applicable (e.g., Poulin and Krishnan, 1995,1996a,b; Poulin and Thiel, 2000). At the present 32 time, there is no validated animal-replacement approach for predicting association constants for 6/29/05 3-11 DRAFT—DO NOT CITE OR QUOTE ------- 1 blood or tissue protein binding of organic chemicals (Poulin and Krishnan, 1996b). The 2 biologically based algorithms as such are useful in providing initial estimates of tissue:blood 3 partition coefficients solely based on the consideration of solubility in water and lipid contents of 4 tissues and blood. A number of other empirical or semiempirical methods relating molecular 5 structure or physicochemical characteristics to tissue:blood and blood:air partition coefficients of 6 chemicals are also available (Payne and Kenney, 2002; Beliveau and Krishnan, 2003) (Appendix 7 1). Their use is acceptable, as long as the qualitative and quantitative aspects of structural 8 features and physicochemical characteristics of the new chemical are within those used in 9 calibrating the algorithm. 10 Partition coefficients required for PBPK models should be obtained using in vitro 11 methods, in vivo data obtained at steady-state, or theoretical algorithms within the boundary of 12 valid application. 13 14 3.4.3. Biochemical Parameters 15 Absorption rates, metabolic parameters (e.g., first-order or second-order rate constants, 16 maximal velocity, and Michaelis affinity constant) and tissue diffusion constants (for describing 17 diffusion-limited uptake) required for PBPK modeling can confidently be determined using in 18 vivo studies. For estimating these parameters, pharmacokinetic data (e.g., time course of tissue 19 or blood concentration of parent chemicals, urinary metabolite levels) obtained following a 20 single bolus dose or infusion may be used. For volatile organic chemicals, the use of exhaled 21 breath and gas uptake studies has frequently been adopted with success (Filser and Bolt, 1979, 22 1981; Andersen et al., 1980; Gargas et al., 1986). The rate constants of chemical reaction with 23 hemoglobin and tissue proteins determined in vitro or in vivo have been incorporated into the 24 PBPK model to make predictions of these phenomena in vivo (e.g., Krishnan et al., 1992). The 25 use of in vivo data for parameter estimation is acceptable, as long as the same data set is not used 26 for model evaluation purposes as well. The use of Bayesian approach is likely to enhance the 27 precision of parameter estimations from in vivo data (e.g., Vicini et al., 1999). 28 Receptor binding and DNA-binding properties of chemicals have been successfully 29 described with PBPK models on the basis of in vitro-derived data (Terasaki et al., 1984; Farris et 30 al., 1988). The appropriateness of the usefulness of in vitro systems (e.g., freshly isolated 31 hepatocytes, microsomes, post-mitochondrial fractions, cytosols) to provide metabolic constants 32 for incorporation into PBPK models continues to be an active area of investigation. These data 6/29/05 3-12 DRAFT—DO NOT CITE OR QUOTE ------- 1 may be applicable to modeling using the parallelogram approach. For example, chemical- 2 specific in vitro metabolic data from cultured hepatocytes can be scaled to represent in vivo liver 3 clearance using in vitro data such as estimates of the number of hepatocytes present per one gram 4 of liver tissue and the average liver weight (in grams) of the species and age group of interest. 5 In vitro data for humans is then extrapolated to in vivo by assuming that the same relationship 6 that successfully describes the in vitro to in vivo relationship in animals effectively converts the 7 human in vitro data to the in vivo situation. Even though there are several examples of 8 successful application based on appropriate in vitro-in vivo scaring methods (Reitz et al., 1989, 9 1996a; Gearhart et al., 1990; Hwang et al., 1996; De Jongh and Blaauboer, 1996,1997; Iwatsubo 10 et al., 1996; Kedderis and Held, 1996; Mortensen et al., 1997; Mortensen and Nilsen, 1998; Cole 11 et al., 2001; Hissink et al., 2002; Lipscomb et al., 1998, 2003), the extrapolation of in vitro data 12 to intact animal is not clear in all cases (e.g., Haddad et al., 1997,1998). But the in vitro studies 13 are particularly useful for evaluating the extent of metabolism in target tissues, characterizing 14 interindividual differences in metabolism, and conducting animal-human extrapolation of 15 metabolism constants based on a parallelogram approach (Andersen et al., 1991; Reitz et al., 16 1996b; Ploeman et al., 1997; Thrall et al., 2000; Kedderis and Lipscomb, 2001). 17 18 Biochemical parameters for PBPK models can be estimated using in vivo data or on the 19 basis of adequate scaling of in vitro data. 20 21 3.5. COMPUTER IMPLEMENTATION 22 Most of the PBPK models require the use of numerical simulation methods because they 23 contain differential equations and descriptions of nonlinear processes. Therefore, the PBPK 24 model equations are written along with the integration algorithms and solved using programming 25 languages, simulation software, or spreadsheets. Simulation languages and commercially 26 available packages (Table 3-7) routinely make use of integration algorithms to obtain numerical 27 solution to differential equations (Menzel et al., 1987; Burmaster and Murray, 1997; Easterling et 28 al., 2000, Reddy et al., 2003) such that there is no need either for the modeler or the risk assessor 29 to evaluate this aspect. However, if a programming language (FORTRAN, BASIC) or 30 spreadsheet (Lotus 1-2-3, QuattroPro, Microsoft Excel) is used for modeling, then the modeler 31 should write the codes for an appropriate numerical integration algorithm (e.g., Euler, Gear, 32 Runge-Kutta routines; predictor-corrector methods). In such cases, the integration algorithm as 6/29/05 3-13 DRAFT—DO NOT CITE OR QUOTE ------- Table 3-7. Examples of simulation software used for PBPK modeling Software Fortran compiler with IMSL library packages, C, Pascal, Basic ACSL, ACSL-Tox,or acslXtreme (Advance Continuous Simulation Language) SimuSolv Matlab Microsoft Excel ScoP (Simulation Control Program) Stella Developer/vendor Many vendors sell different compiler packages available on the market. The Aegis Technologies Group, Inc., Huntsville, AL Dow Chemical Company, Midland, MI (no longer distributed outside the company) The MathWorks, Natick, MA Microsoft Corporation, Redmond, WA Simulation Resources, Inc., Redlands, CA Isee Systems, Lebanon, NH (formerly High Performance Systems Inc.) f Salient features Machine language compiler packages that require certain knowledge of computer programming; models can be customized to simulate specific condition The most commonly used for PBPK modeling in the toxicology community. Language designed for modeling and evaluating the performance of continuous systems described by time- dependent, nonlinear differential equations. Makes use of ACSL language to write the dynamic nonlinear systems that are translated into FORTRAN at run time Mathematical software with matrix-related computations, numerical integration algorithms capable of solving systems of ordinary differential equations, and graphical nonlinear simulation (Simulink) Neither translation of the model nor the compilation into a program is required, but integration algorithm and interval should be specified by the user. An interactive control program for constructing models; when used with a C compiler, SCoP greatly simplifies the construction of a simulation program Macintosh, interactive graphical user interface software; enables the user to generate models with diagrams, where a minimal knowledge of computer programming is required Examples of application Hoang (1 995); Karba et al. (1990) Ramsey and Andersen (1984); Thomas etal. (1996); Dong (1994) Rey and Havranek (1996) Easterling et al. (2000) Johanson and Naslund (1 988); Haddad et al. (1 996) Menzel etal. (1987) Hoang (1995) 6/29/05 3-14 DRAFT—DO NOT CITE OR QUOTE ------- 1 2 Table 3-7. Examples of simulation software used for PBPK modeling (continued) Software Mathematica Berkely Madonna SONCHES (Simulation of Nonlinear Complex Hierarchical Ecological Systems) CMATRIX BASICA AVS (Application Visualization System) MCSim Developer/vendor Wolfram Research, Inc., Champaign, IL Robert Macey and George Oster, University of California at Berkeley, CA Central institute of Cybernetics and Information Processes, Academy of Sciences of GDR, Berlin Robert Ball and Sorell L. Schwartz, Georgetown University, Washington, DC California Department of Pesticide Regulation, Sacramento, CA Advanced Visual Systems, Inc., Waltham, MA Drs. Bois and Maszle Salient features Mathematical software with matrix-related computations; numerical integration algorithms capable of solving systems of ordinary differential equations This program is a general- purpose differential equation solver. Developed on the Berkeley campus under the sponsorship of National Science Foundation and the National Institutes of Health. It is currently used by academic and commercial institutions for constructing mathematical models for research and teaching. A computer system where connections between various data libraries in the preparation and post- processing of simulation are executed by macro commands A system that allows the user to create compartmental models based on personal biological knowledge, leaving the construction and numerical solution of the differential equations to the software Numerical integration algorithms developed by the Department for PBPK modeling A visualization software package capable of importing processed resonance images and combining the use of ACSL to create three-dimensional representations of the PBPK of a chemical in an organism This software, facilitates the conduct of Bayesian analysis with PBPK models but has no graphical interphase. Examples of application Burmaster and Murray (1997) Reddy et al. (2003) Wunscheretal. (1991) Ball and Schwartz (1994) Dong (1994) Nichols etal. (1994) Jonsson and Johanson (2003) 6/29/05 3-15 DRAFT—DO NOT CITE OR QUOTE ------- 1 well as the integration interval used should be specified (e.g., Blancato and Saleh, 1994; Haddad 2 et al., 1996). 3 The modeler should also be aware of the optimization routine offered by software 4 packages, particularly if parameters are to be estimated from experimental data by statistical 5 optimization (Holmes et al., 2000). The personal and portable computers marketed today 6 possess the acceptable speed, disk space, and run time memory required for PBPK modeling and 7 parameter optimization; therefore, this aspect needs no formal evaluation. 8 The accuracy of computational representation of PBPK models is evaluated by 9 "debugging," which refers to the process of error detection in computer programs. The "bugs" in 10 PBPK models written as a program may result either from typing errors or from illogical 11 mathematical statements. To eliminate these errors, it is essential to carefully verify the model 12 codes after entry into the computer. Commercially available simulation software, while 13 converting the model codes written in a source language to machine language, can detect 14 syntax/language errors related to incorrect writing of model codes. However, such error 15 diagnostic features cannot detect errors associated with incorrect mathematical representation of 16 a process written in correct programming language without any typing mistakes. The onus is on 17 the modeler to ensure that the equations are entered correctly using the programming/simulation 18 software, and risk assessors cannot be expected to go through the PBPK model codes to perform 19 a routine error diagnostic check. Such verification should be done by the developer initially and 20 then by individuals not involved in model development, such as peer-reviewers and co-workers 21 (Clark et al., 2004). 22 The computer implementation of PBPK model need not be evaluated by risk assessors if a 23 highly reputable commercial or open source simulation software has been used. The onus is on 24 the model developer to ensure that the computer implementation of the PBPK model is free from 25 errors. When the modeler writes his/her own program, the appropriateness of the integration 26 algorithm and integration intervals should be justified; similar concerns would exist initially for 27 newly developed commercial or open source software. 28 29 3.6. MODEL EVALUATION 30 The purpose of model evaluation is to assess the adequacy of the model and its 31 parameters to consistently describe the available data sets representing the pharmacokinetics of a 32 chemical. A model that has been adequately evaluated would not only capture the critical 6/29/05 3-16 DRAFT—DO NOT CITE OR QUOTE ------- 1 determinants of the pharmacokinetics of a chemical but also characterize the elements of 2 uncertainty associated with the parameters. 3 The first step to ensure validity of the PBPK model is to check on the mass balance of the 4 chemical in the whole body. The model should not create or destroy matter. Basically, when all 5 the tissue levels, blood levels, and any excretion or metabolic transformation levels are added 6 they should sum up to the same exposure dose level used in the model. Although trivial, this 7 check is important in understanding the sources in the model that would falsely increase or 8 decrease the level of the chemical in tissues. Another initial evaluation step is to check the model 9 behavior when exposure is set to zero. The zero check is necessary to make sure that the model 10 can represent steady-state levels of the chemical and to make sure that this level does not change 11 with time in the absence of exposure. 12 The PBPK model evaluation to date has not been done rigorously from a statistical 13 perspective. The most common approach to model evaluation has involved visual inspection of 14 the plots of model predictions (usually continuous and represented by solid lines) with 15 experimental values (usually discrete and represented by symbols) against a common 16 independent variable (usually time). The rationale underlying this qualitative approach to model 17 evaluation is that the greater the commonality between the predicted and experimental data, the 18 greater the confidence in the model structure and parameters. The correspondence between 19 predictions and experimental data should be not only at the level of numbers (e.g., blood 20 concentration values) but also at the level of the profile (i.e., bumps and valleys in the 21 pharmacokinetic curve). Because PBPK modeling is not a fitting exercise, model simulations 22 are not expected to touch upon each experimental data point. Rather, the shape of the simulated 23 curve should correspond to that of the experimental data, and the simulated data should 24 preferably be within two standard deviations of the mean of experimental values, even though 25 this is not a rule of thumb used by the modeling community. 26 Figure 3-2 shows several examples of visual evaluation of the adequacy of PBPK models. 27 The models used in cases A and C would be considered adequate because they simulate the 28 behavior of the experimental data even though they do not accurately simulate every single 29 experimental data. On the other hand, models used in cases B and D would not be considered 30 adequate, and further work would be required to refine the models, either because they do not 31 simulate the profile appropriately (B) or the model simulations are outside two standard 32 deviations of the experimental data(D). 33 6/29/05 3-17 DRAFT—DO NOT CITE OR QUOTE ------- 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 A _ I — I—. c I ^ — V, i 1 - 1 °-1-f \ * 1 \^ 5 ai~ Om . , n m Time B Time C D 1 1 1 " . r T-\ •I /"^~^ V I 2 / ^v £5 1 ° 1 \ 1 °-1 ' 0 ' * \»_ o O \v^ O OrH — i n 0*1 - I 5 , f \^ Time Time Figure 3-2. Comparison of PBPK model simulations (solid lines) with experimental data (symbols). The above approach to model evaluation says nothing about the adequacy of the model structure or parameters. It only reflects an individual's judgment of how closely the model predicts the observed behaviour. Evaluating the adequacy of model structure and equations is fairly straightforward when compared with the evaluation of the model parameter values. For example, inadequacies in PBPK model structures can be inferred simply by observing the simulated and experimental pharmacokinetic profiles. Figure 3-2 (B) depicts a model whose structure is inadequate because the simulated profile is not consistent with the experimental data for the chemical. As a caveat, it is possible that the lack of fit in Figure 3-2 (B) is due to a problem with the model parameters, not the structure. If the model cannot fit the 6/29/05 3-18 DRAFT—DO NOT CITE OR QUOTE ------- 1 pharmacokinetic profiles for any realistic parameter values, or only using values that are 2 inconsistent with other data, then one can reasonably conclude that the structure is inadequate. 3 This evaluation of model structure provides the developer an opportunity to think about 4 the need for additional compartments, critical determinants of disposition, or different 5 quantitative descriptions of the phenomena, and to improve the capability of the model, 6 accordingly. A useful way of comparing the experimental and simulated data is to plot the 7 residuals (i.e., difference between experimental and simulated data) as a function of time or as a 8 function of various controllable variables. If more than one model fits the experimental data 9 equally well, new experiments may be designed to identify the model that more accurately 10 predicts the other attributes of the biological system (Kohn, 1995). 11 Prior to the discussion about evaluating model parameters, it is important to correct a 12 common misunderstanding about what a "validated model" means. A model that has been 13 calibrated against one dataset, and that adequately simulates a different data set, can be said to be 14 "validated," but it is only validated to the extent to which those two dataset accurately represent 15 the larger population, not in any global sense independent from the data used to develop the 16 model. PBPK models are used to extrapolate to other exposure conditions or dosing regimens, 17 but here again, only to the extent that the data used to calibrate and test the model are of 18 sufficient quality to support the extrapolations. To avoid giving the impression that a model is 19 "validated" to predict outcomes for which it has not been adequately tested, many PBPK 20 modelers today prefer to use the terms "calibrated" model (accompanied by a description of the 21 data used to calibrate the model) and the "predictive capability" or utility of the model, again 22 specifying the data that the model was adequately able to simulate. In this sense, the more quality 23 data available to calibrate and test the model, the more confidence one can place in the model's 24 predictive capability. PBPK modelers are in the process of developing better qualitative and 25 quantitative descriptors for both models and the data used to develop models, to assist risk 26 assessors in their evaluation and application of PBPK models. 27 The adequacy of model parameter values may be evaluated in different ways; no single 28 method has been accepted or endorsed by the modeling or regulatory community as yet. 29 Statistical methods required for evaluating the adequacy of model parameters are based on 30 comparison of simulations with experimental data, and depend on whether the objective is to 31 perform internal evaluation (in which all model parameters are estimated from the same data 32 set), external evaluation (in which different datasets are used for model calibration and testing 6/29/05 3-19 DRAFT—DO NOT CITE OR QUOTE ------- 1 the predictive capability of the model) or serai-external evaluation (in which some of the model 2 parameters are based on the data set). Although no systematic research effort or guidance is 3 available in this regard, there is much interest in developing consistent and acceptable evaluation 4 methods, and progress is being made. 5 For external evaluations, none of the classical tests (t, Mann-Whitney, two-sample x2, 6 two-sample Kolmogrov, etc.) that determine whether the underlying distributions of the two 7 datasets are similar is applicable, since the output processes of almost all real world systems and 8 simulations are non-stationary and autocorrelated. Furthermore, there is a question of whether 9 the use of statistical hypothesis tests is even appropriate. Since the model is only an 10 approximation of the actual system, a null hypothesis that the system and model are the same is 11 clearly false. The more appropriate question is to ask whether or not the differences between the 12 system and the model are significant enough to affect conclusions derived from the model. In 13 this regard, Haddad et al. (1995) screened various statistical procedures (correlation, regression, 14 confidence interval approach, lack fit F test, univariate analysis of variance, and multivariate 15 analysis of variance) for their potential usefulness in evaluating the degree of agreement between 16 PBPK model simulations and experimental data According to these authors, the multivariate 17 analysis of variance represents the most appropriate statistical test for the purpose of comparing 18 . PBPK model predictions with experimental data. For now, however, the visual inspection is the 19 most frequently used approach, if not the best approach, to the external evaluation of the 20 performance of PBPK models. 21 One approach for determining if the level of complexity (number of parameters) in a 22 model is justified by the data is to use a nested modeling approach, where the model is reduced 23 to a simpler (nested) model when one or more parameters are set to zero. The increase in 24 goodness-of-fit obtained by allowing those parameters to be non-zero can then be evaluated 25 statistically using a x2 statistic, to determine if the additional degrees of freedom afforded by 26 those parameters are justified (Collins et al., 1999). 27 There is increasing concern about the relevance and usefulness of external evaluation in 28 PBPK modeling, particularly as it relates to humans. External evaluation requires that some of 29 the available pharmacokinetic data not be used during model calibration phase, but kept for 30 evaluating the performance of the model. Not everyone is in agreement with such an approach. 31 Some people believe that all the data used for model evaluation should be used to improve the 32 parameter estimates, so that no data are "wasted" towards that end. Such an iterative approach to 6/29/05 3-20 DRAFT—DO NOT CITE OR QUOTE ------- 1 model evaluation and calibration maximizes the use of the available data. This is particularly an 2 issue of relevance to human data since the actual parameters for each individual in a population 3 might be different such that a model with a single set of parameters may not be reasonably 4 expected to simulate the observed kinetics in all individuals. Therefore, the process of modeling 5 should not only take into account the existing information on parameters, but also be able to 6 accommodate the new information based on fits to additional datasets. In this context, Bayesian 7 analysis is being increasingly explored for use in PBPK modeling (Bois et al., ZOOOa, b; Jonsson 8 and Johanson 2001, 2002). In the Bayesian approach, the prior information on parameters is 9 updated with new pharmacokinetic data such that the resulting posterior estimates consistently 10 describe all data, and support better characterization of the uncertainty and distribution in the 11 parameter values. Cross validation is another potentially useful approach in this regard (Keys et 12 al., 2003). As used in structure-activity relationship modeling arena (Beliveau et al., 2003), cross 13 validation involves using all the available data sets by repeated sub-sampling. This type of a 14 leave-one-out cross validation allows the use of available data both for estimation and evaluation 15 of model parameters. It is likely that no single approach will be sufficient or applicable in all 16 contexts. Each of these approaches has its merits and limitations, and their applicability for 17 PBPK model evaluation depends upon the purpose of the model and data availability. 18 PBPK models intended for use in risk assessment should be evaluated to ensure that they 19 provide simulations of pharmacokinetic profile consistent with the experimental data and that 20 the parameters (point estimates, range of values, or distributions) are appropriate for the 21 intended application. 22 23 3.7. SENSITIVITY, UNCERTAINTY, AND VARIABILITY ANALYSES 24 The variability, uncertainty, and sensitivity of parameters constituting the PBPK models 25 may need to be evaluated, depending on the intended application(s) of the model. 26 27 3.7.1. Sensitivity Analysis 28 Sensitivity analysis deals with the quantitative evaluation of how parameters or input 29 functions influence the outcome. Such an analysis conducted in the context of risk assessment 30 would identify how each parameter influences the risk estimate and which specific one(s) 31 drive(s) the risk estimate (U.S. EPA, 2001). By identifying the most sensitive parameters, 32 sensitivity analysis facilitates focused use of resources for uncertainty and variability analyses. 6/29/05 3-21 DRAFT—DO NOT CITE OR QUOTE ------- 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 Sensitivity analysis in the context of PBPK modeling involves determining the magnitude of change in pharmacokinetic simulations for a defined change in the input parameters. The results of sensitivity analysis are the sensitivity ratios that correspond to the ratio of change in simulation output (e.g., tissue dose) to change in parameter value. Figure 3-3 depicts the sensitivity ratios associated with some input parameters of a PBPK model. The greater the absolute value of the sensitivity ratio, the more important the parameter. In this example, the sensitivity ratio for breathing rate is the highest of all input parameters, indicating that it is the most sensitive PBPK model parameter with respect to the dose metric (i.e., parent chemical AUC in target tissue). In this case, a sensitivity ratio of 2 associated with the breathing rate signifies that 1% change in the numerical value of this parameter will lead to a 2% change in the dose metric (Figure 3-3). 3 - 2 - o •S3 (D (2 2? . ;5 1 - 0) W -1 - ' :jjjlj!|:;:;:|:|S :W:$:*x:x::: 0 0) .t 0) X ra ^ c ro c ra if I 12 |i > o 2 0) m CD Figure 3-3. Sensitivity ratios associated with certain input parameters of a PBPK model. Sensitivity analysis is informative for identifying the key parameters that are likely to affect the performance of the model and output such that efficient resource allocation for further research to enhance confidence in the parameter estimates can be done (e.g., Bois et al., 1991; Hetrick et al., 1991; Clewell et al., 1994). A sensitivity analysis is not obligatory for a PBPK 6/29/05 3-22 DRAFT—DO NOT CITE OR QUOTE ------- 1 model to be acceptable for risk assessment applications, but it greatly strengthens the credibility 2 of the model and guides resource allocation for risk assessment-oriented research. 3 One caveat when conducting individual parameter sensitivity analyses (such as described 4 above) is that they only show the sensitivity of the model predictions to a change in the single 5 parameter when all other parameters are held constant. For example, consider the sensitivity to 6 breathing rate depicted in Figure 3-3. What would have happened if one had known, before 7 starting the modeling process, that the breathing rate was 20% higher than the default value 8 actually used? Would the predicted dose metric have then turned out to be 40% higher? Only if 9 none of the other parameters were calibrated to the model data during the modeling process. If 10 in each case (default breathing rate vs. 40% higher) one had started with that value of the 11 breathing rate, and then calibrated the Vmax and other parameters to the data, the result would be 12 different values of Vmax, etc., that would compensate to some extent for the change in Vmax. As a 13 simple analogy, consider the fit of a straight line to some data, where the intercept is a "known" 14 parameter and the slope is fitted. After fitting the line one might determine that the "fit" is very 15 sensitive to the intercept by showing that if the intercept is changed while holding the slope 16 constant (i.e., standard sensitivity analysis), the value of the line equation changes a lot. But if 17 one had started with a larger value of the intercept at the beginning of the "modeling" process, 18 fitting the line to the data would have resulted in a lower value for the slope, such that the value 19 of the line equation would not change as much as when only the intercept is increased. In short, 20 the parameter estimation process leads to certain correlations between the values of parameters 21 that are fixed as inputs and those that are fitted. Thus, standard sensitivity analysis, while very 22 informative about the importance of individual parameters, over-estimates the actual impact of 23 changes in individual parameters because it does not account for correlations. In the case of 24 Figure 3-3, it might be that starting with a different breathing rate and then calibrating Vmax, etc., 25 would have yielded almost identical values for the dose metric, and that the overall modeling 26 process is insensitive to breathing rate, even though the model predictions are sensitive to 27 changes in breathing rate when none of the other parameters are changed. 28 29 3.7.2. Variability Analysis 30 The focus of a variability analysis is to evaluate the range of values for a parameter 31 expected to be present in individuals of a population and its impact on tissue dose simulations. 32 PBPK models accounting for individual-specific parameters (e.g., enzyme levels, tissue volumes, 6/29/05 3-23 DRAFT—DO NOT CITE OR QUOTE ------- 1 body weights, workload) may be constructed to simulate tissue dose variability in populations 2 (e.g., Sato et al., 1991; Dankovic and Bailer, 1994). Alternatively, PBPK models for an average 3 individual representing specific subgroups of populations (e.g., adult women, pregnant women, 4 lactating women, children) may be constructed by accounting for subgroup-specific 5 physiological, biochemical, and physicochemical parameters (e.g., Fisher et al., 1997; Krishnan 6 and Andersen, 1998; Corley et al., 2003; Price et al., 2003b). These analyses, however, would 7 not provide the probability or likelihood of a particular output for a population. 8 With information such as the tissue dose corresponding to the 95th percentile and 50th 9 percentile, the magnitude of the interindividual variability factor can be computed. For this 10 purpose, Monte Carlo simulation approaches based on prior distributions of input parameters 11 (physiological parameters, enzyme content/activity with or without the consideration of 12 polymorphism) have frequently been used (Thomas et al., 1996; El-Masri et al., 1999; Bogaards 13 et al., 2001; Haber et al., 2002; Lipscomb and Kedderis, 2002; Timchalk et al., 2002; Lipscomb 14 et al., 2003). This method involves repeated computations using inputs selected at random from 15 statistical distributions for each parameter in order to provide a statistical distribution of the 16 output, i.e., dose metric (Figure 3-4). The Monte Carlo approach to variability analysis has also 17 helped in evaluating the net impact of the variability of critical biochemical and physiological 18 parameters (e.g., Portier and Kaplan, 1989). 19 When conducting variability analysis, it is important that correlations in model 20 parameters be included in the evaluation. For example, cardiac output (Qc) and breathing rate 21 (QA) are expected to vary in proportion to each other, so using independent distributions that 22 might give a very high value of Qc with a very low value of QA would be unrealistic. On the 23 other hand, one could consider the distributions of Qc and the distribution of the fAc = QA/Q, and 24 multiply the value selected from the fAc distribution by the value selected from the Qc 25 distribution to obtain the value of QA to be used. 26 More recent research has indicated the potential usefulness of Bayesian framework based 27 on Markov Chain Monte Carlo simulation (Jonsson and Johanson, 2001, 2002). This method 28 combines prior knowledge about parameters and data from new experimental studies to generate 29 posterior parameter distributions, which in turn are used in Monte Carlo simulations for 30 conducting variability and uncertainty analyses (Bois, 1999; Johanson et al., 1999; Bernillon and 31 Bois, 2000). 32 6/29/05 3-24 DRAFT—DO NOT CITE OR QUOTE ------- 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 Metab $jbj|fi zy m es " " " f ff f Partition fcospcients Monte Carlo Simulation Time INPUT DISTRIBUTIONS COMPUTER PROGRAM SIMULATIONS Figure 3-4. Illustration of the use of the Monte Carlo approach for simulating the distribution of internal concentration versus time on the basis of population distributions of PBPK model parameters. Conducting a variability analysis with a PBPK model is not a prerequisite for its use in risk assessment applications. The assessment of the impact of parameter variability on tissue dose, however, is a prerequisite for a PBPK model intended for use in estimating the interindividual variability factor (pharmacokinetic component). 3.7.3. Uncertainty Analysis Uncertainty analysis in the context of PBPK modeling refers to the evaluation of the impact of the lack of precise knowledge about the numerical value of a parameter or model structure itself on dose metric simulations. The uncertainty regarding model structure and parameter values may contribute to uncertain predictions of dose metrics, particularly for low- dose exposure situations (Hattis et al., 1990). The uncertainty analysis is particularly useful when a PBPK model does not adequately simulate the experimental data. Such a situation may arise due to either lack of precise estimates of parameter values or inadequacy of the model structure chosen for the study. In these cases, either a quantitative uncertainty analysis or model- directed mechanistic studies should help improve the predictive ability and robustness of the PBPK model (Clewell and Andersen, 1997; Haddad et al., 1998). 6/29/05 3-25 DRAFT—DO NOT CITE OR QUOTE ------- 1 Quantitative uncertainty analyses of PBPK model have frequently been conducted using 2 either a Monte Carlo approach or Bayesian framework based on Markov Chain Monte Carlo 3 simulations with respect to specific endpoints (e.g., amount metabolized, tissue concentration of 4 parent chemical at a specific time, cancer estimates) (e.g., Farrar et al., 1989; Krewski et al., 5 1995; Gelman et al., 1996; Elder, 1999). Alternatively, a stochastic response surface method or a 6 fuzzy simulation approach may be used for uncertainty analysis with PBPK models (Isukapalli et 7 al., 1998; Nesterov, 2001). The latter method is particularly useful when statistical distributions 8 of parameters cannot be defined reliably and only semi quantitative, qualitative, and vague 9 information is available. 10 If there is a lack of confidence regarding the numerical value of a parameter (e.g., 11 imprecision due to the method used for parameter estimation), a quantitative analysis of the 12 uncertainty associated with parameter(s) of the PBPK model should be conducted. Such analysis 13 will provide an indication of the impact on tissue dose simulations of the imprecision or lack of 14 accurate knowledge of a parameter value. However, if a PBPK model has been evaluated using 15 various sets of data (e.g., species, dose levels, routes), then the benefits of uncertainty analysis 16 are limited. Particularly, if the model with its set of parameters has been shown to predict 17 adequately the kinetics of a chemical following various dose levels administered by more than 18 one route in more than one species, conducting a quantitative uncertainty analysis cannot help 19 improve any further the closeness of model predictions to experimental data. In other words, the 20 improvement of the precision of parameter estimates, after a certain level, may not help improve 21 the predictive ability of the model. Models that have been adequately evaluated may therefore 22 be used for risk assessment applications without detailed uncertainty analysis. Where possible 23 and relevant, uncertainty analysis should be performed to strengthen the credibility of the PBPK 24 model and guide resource allocation to risk assessment-oriented research. 25 The conduct of sensitivity, uncertainty, and variability analyses should be based on 26 acceptable statistical methods. EPA has published guiding principles for Monte Carlo analysis 27 (U.S. EPA, 1997), but there is no such guidance for Bayesian and Markov Chain Monte Carlo 28 methods. When using these methods, care should be taken to ensure that the resulting PBPK 29 model simulations respect the following basic conditions: 30 • The numerical values of physiological parameters (representing prior or posterior 31 distributions) are within known, plausible limits; 6/29/05 3-26 DRAFT—DO NOT CITE OR QUOTE ------- 1 • The sum of tissue volumes is lower than the body weight; 2 • The sum of tissue blood flows is equal to cardiac output; 3 • The mass balance is respected (chemical absorbed = chemical in body + chemical 4 eliminated); and 5 • The covariant nature of the parameters is appropriately respected (e.g., the person 6 with lowest breathing rate should not be the one receiving the highest cardiac output) 7 8 While taking advantage of the sophisticated statistical approaches, it is important to 9 ensure that the resulting model and parameters are within plausible range or representative of the 10 reality. 11 Sensitivity, uncertainty, and variability analyses can help improve the credibility of 12 PBPK models as well as prioritize research needs to improve the model for risk assessment 13 applications. However, such analyses may not be required for all PBPK models intended for 14 risk assessment applications. 15 16 3.8. DEVELOPING PBPK MODELS FOR USE IN RISK ASSESSMENT: 17 STRATEGIES FOR DEALING WITH DATA-POOR SITUATIONS 18 3.8.1. Minimal Data Needs for Constructing PBPK Models 19 When an adequately evaluated PBPK model is not available for the species, life stage, 20 and route relevant to a risk assessment application, significant resources may be needed to 21 develop such a model, depending on the chemical, the availability of prior information, and the 22 complexity of disposition mechanisms being modeled. The minimal data required for 23 developing such models for a chemical in any given species are: 24 • Partition coefficients, 25 • Biochemical constants, 26 • Route-specific absorption parameters, and 27 • In vivo pharmacokinetic data for model evaluation. 28 29 As outlined in this chapter, the partition coefficients required for PBPK modeling may be 30 estimated using the theoretical algorithms found in the literature (Appendix 1). Their use, 31 however, should be limited to the domain of validity and the families of chemicals for which 6/29/05 3-27 DRAFT—DO NOT CITE OR QUOTE ------- 1 such algorithms have been developed and validated. Biochemical constants such as metabolism 2 rates may be obtained using in vitro systems. Other biochemical parameters, such as binding 3 association constants and renal clearance, may be required, depending on the chemical. 4 Additionally, route-specific absorption parameters such as the rate of oral absorption and the skin 5 permeability constant are required for describing oral absorption and dermal absorption, 6 respectively. Of these, the skin permeability coefficient can be obtained using available 7 quantitative structure-activity relationships (QSARs) (Appendix 1). However, such absorption 8 parameters are not required for simulating intravenous administration and inhalation exposures. 9 Finally, some in vivo pharmacokinetic data (at a minimum blood concentration time-course data 10 at two dose levels) are required for evaluating the PBPK model for a particular route of 11 exposure. 12 The minimal data set identified above should be available for the species used in the 13 critical study. Human models, however, may be constructed with knowledge of species-specific 14 blood solubility/binding characteristics. Other model parameters, including metabolism rates, 15 may be either scaled or kept species-invariant according to the current state of knowledge 16 (Section 4.5). The availability of the data set for external evaluation in humans, of course, may 17 be a limiting factor. In such cases surrogate data sets may be used for model evaluation 18 purposes. 19 20 3.8.2. Surrogate Data for Interspecies and Interchemical Extrapolations 21 In the absence of human data for model evaluation purposes, a parallelogram approach 22 based on surrogate data has been used successfully. This approach uses two data sets (e.g. one 23 demonstrating the relationship between in vitro and in vivo findings in a test species and the 24 other demonstrating the relationship between in vitro human and in vitro test species findings) in 25 order to predict the in vivo effects in humans. Accordingly, if human data either cannot be 26 collected or is not available for a chemical of interest, it suffices to evaluate a related chemical 27 for which such data are available. Jarabek et al. (1994) used this parallelogram approach for 28 model development and interspecies extrapolation of the pharmacokinetics of HCFC-123 (2,2- 29 dichloro-l,l,l-trifluoroethane). In this case, the authors developed rat PBPK models for HCFC- 30 123 as well as a structural analog (halothane) by estimating partition coefficients and metabolic 31 constants. Following the evaluation of the rat PBPK model for each of these chemicals, human 32 models were constructed. The adequacy of the human model for halothane was evaluated using 6/29/05 3-28 DRAFT—DO NOT CITE OR QUOTE ------- 1 available pharmacokinetic data and by structural and metabolic analogy. HCFC-123 was 2 assumed to be able to reasonably simulate the in vivo pharmacokinetics in humans, even though 3 human data were not available for this chemical for model evaluation purposes (Jarabek et al., 4 1994; Williams et al., 1996). This is one practical way of getting around the lack of human data 5 for model evaluation, particularly when external evaluation is intended. 6 To deal with situations of lack of data on PODs for closely related chemicals, a family 7 approach has been suggested. This approach, proposed by Barton et al. (2000), is based on the 8 principle that the acceptable concentrations for related chemicals, particularly metabolites, can 9 be derived using data on the parent chemical. Thus, if the NOAEL for the parent chemical is 10 established, there would also have been internal systemic exposures to its metabolites. By 11 determining the external exposure levels for these compounds that result in the same systemic 12 exposure, the NOAELs for these compounds can be established. The determination of the 13 internal dose and systemic exposures for the parent chemical and metabolites is accomplished 14 using PBPK models, thus facilitating the derivation and establishment of the RfD/RfC with a 15 poor database. 16 QSAR approaches are also available for constructing inhalation PBPK models for 17 volatile organic chemicals in the rat (Beliveau et al., 2003). Accordingly, the contributions of 18 various molecular fragments (CH3, CH2, CH, C, C=C, H, Cl, Br, F, benzene ring, and H in 19 benzene ring) toward the parameters of PBPK models have been determined. With the 20 knowledge of the number of the fragments occurring in a given molecule, the partition 21 coefficients and metabolic constants required for constructing a first-generation PBPK model can 22 be obtained. This QSAR approach is useful to initially develop PBPK models for other 23 chemicals, as long as the number and nature of fragments do not exceed the ones in the 24 calibration set used in the study (Beliveau et al., 2003). 25 Finally, in dealing with data-poor situations, the parameters, algorithms, and references 26 provided in the appendices provide a good starting point. Useful data and equations for PBPK 27 modeling may also be found in reliable electronic resources. For example, Nestorov (1998) 28 outlines a web-based resource for PBPK modeling (http://www.capkr.man.ac.uk). The site 29 provides instant access to resources such as data, methodology, and tools necessary to start a 30 PBPK modeling effort. Another site pioneered by Dr. Loizou (England) will become available 31 in the near future. 6/29/05 3-29 DRAFT—DO NOT CITE OR QUOTE ------- 1 2 3.9. EVALUATION OF PBPK MODELS: SUMMARY 3 The basic criteria for evaluation of PBPK models, as outlined in Sections 3.1 through 3.7, 4 are summarized below. 5 • The structure of a PBPK model intended for risk assessment applications should 6 contain the target organ (or a surrogate tissue). 7 • The MBDEs in PBPK models should have units of quantity per time (e.g., mg/hr) or 8 concentration per time (mg/L-hr"1), and the type of equations chosen to describe 9 ADME should be justified on the basis of hypothetical or known mechanisms of such 10 processes. 11 • The tissue volumes, flow rates, and ventilation:perfusion ratios specified in the model 12 should be within physiological limits. Particularly, the sum total of the tissue 13 volumes should be within the body weight, and the sum total of tissue blood flow 14 rates should equal cardiac output. 15 16 • The power function used for scaling of physiological flows on the basis of body 17 weight should be within 0.67 and 0.75, unless species- or individual-specific data are 18 available. 19 • The power function used for scaling physiological flows and maximal velocity of 20 metabolism on the basis of body weight should be within 0.67 and 0.75. 21 • Maximal velocities of metabolism may also be scaled on the basis of body weight, but 22 measured values for specific enzymes in humans do not generally correlate with body 23 weight, so the choice of whether and how to scale metabolism is at the discretion of 24 the modeler. 25 • Partition coefficients required for PBPK models should be obtained from in vivo data, 26 in vitro data, or theoretical algorithms in the application domain. 27 • Biochemical parameters for PBPK models should be estimated using in vivo data or 28 valid in vitro methods. 29 • The PBPK model should be implemented using commercially available software. If 30 the modeler chooses to write his/her own program, then the appropriateness of the 31 integration algorithm and integration interval should be justified. 32 33 • It is suggested that the PBPK model code be checked for accuracy of units, mass 34 balance, blood flow balance, and behavior at zero dose. A model used in a risk 35 assessment should be accompanied by sufficient documentation to support an 36 independent evaluation and reconstruction of the model and simulation results. A 37 more rigorous verification that may be considered by the risk assessor is to 6/29/05 3-30 DRAFT—DO NOT CITE OR QUOTE ------- 1 independently re-code the model to assure that the documentation is thorough and that 2 there are no "bugs" in the code. 3 • Evaluation of the PBPK model structure and parameters should be conducted to 4 ensure that the model adequately predicts the pharmacokinetic behavior (i.e., bumps 5 and valleys in the concentration vs. time plot) of the chemical and that the parameters 6 (point estimates, range of values or distributions) consistently describe available data. 7 6/29/05 3-31 DRAFT—DO NOT CITE OR QUOTE ------- 1 4. APPLICATION OF PBPK MODELS IN RISK ASSESSMENT 2 3 4.1. CHOOSING PBPK MODELS APPROPRIATE FOR USE IN RISK ASSESSMENT 4 Whether or not a PBPK model was initially intended for risk assessment purposes, it can 5 be useful for risk assessment if it permits simulations of the tissue and blood concentrations of 6 the toxic moiety (parent chemical or metabolite) associated with the animal toxicity or human 7 epidemiological study serving as the basis for the derivation of health protective values (e.g., 8 RfC, RfD, cancer slope factors). Specifically, the model should be able to simulate the dose 9 metrics in the test species as well as in humans for the exposure route and exposure scenario of 10 relevance. For use in risk assessment, a PBPK model: 11 • Should have been developed or calibrated for species and life stage of relevance to the 12 risk assessment, 13 • Must have been peer-reviewed and evaluated for its structure and parameters, 14 • Should consist of parameters essential for simulating uptake via routes associated 15 with human exposures as well as the critical study chosen for the assessment, and 16 • Should be able to provide predictions of the time-course of concentration of toxic 17 moiety (parent chemical or metabolite) in the target organ or a surrogate 18 compartment. 19 20 Figure 4-1 depicts how the above criteria can be applied for selecting appropriate PBPK models. 21 Basically, a peer-reviewed PBPK model for the relevant species and life stage consisting of 22 parameters for simulating relevant routes of exposure and potentially relevant dose metrics is 23 appropriate for use in risk assessment. 24 The first criterion, though appearing self-evident, is quite fundamental, because the 25 models available in the literature sometimes may not have been developed for the specific life 26 stage and species used in the critical lexicological study forming the basis of a risk assessment. 27 For example, PBPK models for methanol have been developed in rats, monkeys, and humans 28 (Horton et al., 1992; Rogers et al., 1993; Bouchard et al., 2001), yet the critical study appears to 29 be a developmental toxicity study in mice, although a newer two-generation study in rats may 30 also be important (Clark et al., 2004). When the PBPK model has not been developed for the life 31 stage or species used in the study forming the basis of the POD for an assessment, additional 32 6/29/05 4-1 DRAFT—DO NOT CITE OR QUOTE ------- Is the PBPK model available for the test species and humans? Model development Are the parameters for simulating relevant routes available? Has the model been evaluated and eer-reviewed? Does the model simulate dose metrics of relevance to risk assessment? Undertake evaluation and peer review Use in risk assessment Use for mechanistic studies & investigation of mode of action 2 3 4 Figure 4-1. Decision tree for selecting PBPK models appropriate for use in risk assessment. 6/29/05 4-2 DRAFT—DO NOT CITE OR QUOTE ------- 1 work may have to be undertaken to 2 resolve the situation before the PBPK 3 model can be used for that particular 4 assessment. 5 The PBPK model for the 6 relevant species and life stage should 7 have been peer-reviewed to determine 8 the adequacy of the structure, 9 parameter estimation methods, and 10 the evaluation efforts. If the PBPK 11 model for a chemical is available in 12 the species and life stage 13 corresponding to the study forming 14 the basis of an assessment but it has 15 not been peer-reviewed, then efforts 16 may be directed towards such a 17 review. Box 4-1 enlists the 18 fundamental questions and issues to 19 be considered while conducting a 20 peer-review evaluation of the 21 adequacy of PBPK models intended 22 for risk assessment applications. 23 Finally, the peer-reviewed model(s) 24 chosen for risk assessment 25 applications should be able to provide 26 simulations of the tissue dose of the 27 toxic moiety or an appropriate dose 28 metric for exposure scenarios and 29 routes associated with the critical 30 study as well as human exposures. 31 Peer-reviewed PBPK models 32 that facilitate the prediction of dose Box 4-1. Questions and issues to be considered while evaluating the adequacy of a PBPK model • Is an appropriate target organ or a surrogate tissue identified as one of the model compartments? • Are the known major sites of storage, transformation, and clearance included in the model structure? • Is the form of equation used to describe chemical uptake justified on the basis of hypothesis of tissue uptake of the chemical? • Have enzymatic processes been described appropriately? (If simulated exposure levels are well below saturation, use of first-order kinetics is acceptable.) • Are the units throughout differential equations consistent (e.g., concentration should not be in mg/ml in one place and \iM in another)? * Are the input parameters related to the characteristics of the host, chemical, or environment? • Is the sum total of compartment volumes within 100% of the body weight? • Do the tissue blood flow rates add up to the cardiac output? • Is the ventilationiperfusion ratio specified in the model within physiological limits? • Is the volume of each tissue compartment within known physiological limits? • Is the approach used to establish partition coefficients valid? • Is the method used for estimating biochemical parameters adequate? • Is the allometric scaling done appropriately? • Is the integration algorithm used in the study known for solving differential equations and dealing with stiff and nonstiff conditions? • Does the shape of the pharmacokinetic curve generated by the model match that obtained experimentally? 6/29/05 4-3 DRAFT—DO NOT CITE OR QUOTE ------- 1 metrics of chemicals for relevant exposure route and life stage of species used in the critical 2 study as well as humans are a prerequisite for their use in risk assessments. 3 4 4.2. EVALUATION OF DOSE METRICS FOR PBPK MODEL-BASED ASSESSMENTS 5 When using PBPK models in risk assessment (RfD, RfC, and unit risk estimates), the 6 basic data needed are: 7 1. POD and critical effect from one or more key studies, 8 2. Peer-reviewed PBPK model for the relevant test species and humans, and 9 3. Dose metric appropriate for the risk assessment. 10 11 The methods and challenges associated with the identification of critical effects and 12 PODs for an assessment remain the same regardless of whether one uses PBPK models or not. 13 The approaches for identifying PODs can be found elsewhere (U.S. EPA, 1994, 2005a). The 14 criteria and issues associated with the selection of PBPK models useful for risk assessment were 15 considered in the previous section. The third data need noted above, i.e., the identification of the 16 appropriate dose metric, is a key aspect determining the use of PBPK models in risk assessment. 17 The dose metric, or the appropriate form of chemical closely associated with the toxicity, 18 varies from chemical to chemical, depending on the MOA and critical effect. The dose metric 19 for PBPK-based risk assessment is chosen following the identification of the potential toxic 20 moiety and evaluation of the relationship with the end point of concern. Analysis of available 21 information on MOA of a chemical along with descriptive toxicity data from various studies, 22 including those using inhibitors or inducers of hepatic metabolism, should be useful in 23 identifying the potential toxic moiety (parent chemical or metabolite). Further, the available data 24 on closely related chemicals may be used to infer the likely toxic moiety. If the induction of 25 toxic effects by a chemical is not altered following treatment with a specific inhibitor of hepatic 26 metabolism, then it would implicate the parent chemical as the toxic moiety. Alternatively, if the 27 induction of the rate of metabolism enhances the extent of toxic effects produced, it would 28 indicate the metabolite as the toxic moiety. Similarly, the toxicity data for various exposure 29 routes and modes of administration may be compared to infer the potential toxic moiety (IPCS, 30 2001). 6/29/05 4-4 DRAFT—DO NOT CITE OR QUOTE ------- 1 After the potential toxic moiety has been identified, the appropriate measure of tissue 2 exposure to the toxic moiety should be chosen (Figure 4-2). For example, the peak concentration 3 has been related to neurotoxic effects of solvents (e.g., Bushnell, 1997; Benignus et al., 1998; 4 Pierce et al., 1998; MacDonald et al., 2002), and concentration of trichloroethylene at the time of 5 testing correlated with effects on behavioural and visual functions (Boyes et al., 2000). Tissue 6 concentrations of TCDD measured during a critical period of gestation have been reported to 7 predict the intensity of developmental responses (Hurst et al., 2000). Gender-specific genotoxic 8 effects of benzene in mice are related to differences in the rate of oxidative metabolism (Kenyon 9 etal., 1996). 10 For chronic effects of chemicals, the integrated concentration of the toxic form of 11 chemical in target tissue over time (i.e., AUC) is often considered a reasonable dose metric 12 (Andersen et al., 1987; Collins, 1987; Voisin et al., 1990; Clewell et al., 2002). For carcinogens 13 producing reactive intermediates, the amount of metabolite produced per unit time and the 14 amount of metabolite in target tissue over a period of time (e.g., mg metabolite/L tissue during 15 24 hours) have been used as dose metrics (Andersen and Dennison, 2001). For developmental 16 effects, the dose surrogate is defined in the context of window of susceptibility for a particular 17 gestational event (e.g., Welsch et al., 1985). Although the AUC and rate of metabolite formation 18 figure among the most commonly investigated dose metrics, other surrogates of tissue exposure 19 may also be appropriate for risk assessment purposes, depending on the chemical and its MOA 20 (e.g., maximal concentration of the toxic moiety, duration and extent of receptor occupancy, 21 macromolecular adduct formation, or depletion of glutathione) (Clewell et al., 2002). Table 4-1 22 lists the dose metrics used successfully in a number of PBPK-based cancer and noncancer risk 23 assessments. 24 When the appropriate dose metric cannot be identified readily, the evaluation of the 25 relationship with the end point of concern should be undertaken with each of the dose metrics in 26 order to identify the one that exhibits the closest or the best association (e.g., Andersen et al., 27 1987; Kirman et al., 2000). This becomes particularly important when there are little or 28 confusing data on the plausible MOA of the chemical. At a minimum, the appropriate dose 29 metric can be identified as the one that demonstrates a consistent relationship with positive and 30 negative responses observed at various dose levels, routes, and scenarios in a given species. In 31 6/29/05 4-5 DRAFT—DO NOT CITE OR QUOTE ------- Potential Toxic Moiety Unknown AUC mg/L tissue x time (Parent chemical or metabolite) Concentration Amount Concentration Amount Rate of production Quantity in tissue \ Quantity in tissue Integrated / Average Cmax (mg/L. Tissue) AUC (mg/L tissue xtime) Cmax (mg/L. Tissue) AUC (mg/L tissue xtime) mg formed per unit time (e.g. mg/hr) mg formed over time specified (e.g. mg/day) Figure 4-2. Examples of measure of tissue exposure to toxic moiety for risk assessment applications. ------- 1 2 Table 4-1. Dose metrics used in PBPK model-based cancer and noncancer risk assessments Chemical Acrylonitrile Bromotrifluoromethane Butoxyethanol (2-) Chloroform Chloropentafluorobenzene Dichlorofluoroethane 1 ,4-Dioxane Ethyl acrylate Ethylene glycol ethers Formaldehye Heptafluoropropane Isopropanol Methoxyacetic acid Methyl chloroform Methyl mercury Methyl methacrylate Endpoint Brain tumors Cardiac sensitization Forestomach lesions and tumors Liver cancer Hepatic effects and kidney tumor Liver toxicity Cardiac sensitization Liver tumors Forestomach tumors Developmental toxicity Cancer Cardiac sensitization Neurobehavioral effects Developmental/ reproductive effects Developmental effects Hepatic effects Neurological effects Nasal lesions Dose metric Peak metabolite concentration in target tissue Concentration of parent chemical at the end of exposure Levels of butoxyethanol/ butoxy acetic acid in forestomach Amount of metabolites covalently bound to biological macromolecules per L liver per day; % cell kill/day Maximal rate of metabolism per unit kidney cortex volume AUC of parent chemical in liver Concentration of parent chemical at the end of exposure Time-weighted average concentration in liver over lifetime Liver AUC Tissue-specific glutathione depletion Peak concentration and average daily AUC of the alkoxyacetic acid (metabolite) in blood DNA-protein crosslinks Concentration of parent chemical at the end of exposure Peak blood concentration AUC of isopropanol and its metabolite (acetone) AUC of parent chemical (gestational day 11) Maximal concentration of parent chemical (g, day 8) Area under the liver concentration vs. time curve Fetal brain concentrations Amount metabolized/time Reference Kirman et al. (2000) Vinegar and Jepson (1996) Poet et al. (2003) Reitzetal. (1990) Meek et al. (2002) Clewell and Jarnot (1994) Vinegar and Jepson (1996) Leung and Paustenbach (1990) Reitzetal. (1990b) Frederick etal. (1992) Sweeney etal. (2001) Casanova et al. (1996); Schlosser et al. (2003) Vinegar and Jepson (1 996) Gentry et al. (2002) Gentry et al. (2002) Clarke etal. (1992, 1993) Welsch etal. (1995) Reitzetal. (1988a) Gearhart etal. (1995) Andersen etal. (1999, 2002) 6/29/05 4-7 DRAFT—DO NOT CITE OR QUOTE ------- 1 2 Table 4-1. Dose metrics used in PBPK model-based cancer and noncancer risk assessments (continued) Chemical Methylene chloride Pentafluoroethane Styrene TCDD Toluene Trichloroethylene Trifluoroiodomethane Vinyl acetate Vinyl chloride Endpoint Cancer Cardiac sensitization Lung tumors (mouse) Biochemical responses Cancer risk Behavioral effects Renal toxicity Neurotoxicity Cancer (liver lung and kidney) Cardiac sensitization Olfactory, degeneration and tumor development Angio sarcoma Dose metric Rate of glutathione metabolites produced/L liver/time Concentration of parent chemical at the end of exposure Steady state concentration of ring oxidation metabolite mediated by CYP2F Body burden Time- weighted receptor occupancy Up/down regulation of receptor occupancy Number of cells induced Brain concentrations at the time of testing Metabolite production/L kidney/day Blood concentration of metabolite (trichloroethanol) Amount metabolized/kg/day; AUC for trichloroacetic acid or dichloroacetic acid in plasma; production of thioacetylating intermediate from dichlorovinylcysteine in kidney Concentration of parent chemical at the end of exposure Intracellular pH change associated with the production of acetic acid; proton concentration in olfactory tissue mg metabolized/L liver; mg metabolite produced/L liver/day Reference Andersen et al. (1987) Vinegar and Jepson (1 996) Cruzen et al. (2002) Kim et al. (2002) Andersen etal. (1993); Portieretal. (1993); Conolly and Andersen (1997) van Asperen et al. (2003) Barton and Clewell (2000) Barton & Clewell (2000) Fisher and Allen (1993); Clewell et al. (2000) Vinegar and Jepson (1 996) Bogdanffy etal. (1999, 2001) Clewell etal. (2001); Reitzetal. (1996b) 3 4 AUC = area under the curve 6/29/05 4-8 DRAFT—DO NOT CITE OR QUOTE ------- 1 other words, the level of the dose metric should be lower for exposure conditions with no effect 2 and higher for toxic exposures, regardless of the route and exposure scenario (Clewell et al., 3 2002). 4 Where there is inadequate basis for giving priority to one dose metric over another, the 5 most conservative one (the dose metric producing the highest risk or lowest acceptable exposure 6 level) should be used in order to be health protective (Clewell et al., 2002). The use of 7 appropriate dose metric helps to reconcile route and species differences in cancer responses, 8 provided there are no pharmacodynamic differences. There has been at least one instance in 9 which PBPK model-derived dose measures could not reconcile rat and mouse kidney tumor data 10 (Smith et al., 1995), indicating the significant role of factors other than the target tissue exposure 11 to toxic moiety. 12 The following section describes the various plausible applications of PBPK models in 13 risk assessment. These applications relate to high-dose to low-dose extrapolation, interspecies 14 extrapolation, intraspecies extrapolation, route-to-route extrapolation, and duration extrapolation 15 as required for RfD derivation, RfC derivation, and cancer risk assessment. 16 17 4.3. EXAMPLES OF THE USE OF PBPK MODELS IN RISK ASSESSMENT 18 4.3.1. High-Dose to Low-Dose Extrapolation 19 PBPK models facilitate high-dose to low-dose extrapolation of tissue dosimetry by 20 accounting for the dose-dependency of relevant processes (e.g., saturable metabolism, enzyme 21 induction, enzyme inactivation, protein binding, and depletion of glutathione reserves) (Clewell 22 and Andersen, 1997). The description of metabolism in PBPK models has frequently included a 23 capacity-limited metabolic process that becomes saturated at high doses. Nonlinearity arising 24 from mechanisms other than saturable metabolism, such as enzyme induction, enzyme 25 inactivation, depletion of glutathione reserves, and binding to macromolecules, have also been 26 described with PBPK models (e.g., Clewell and Andersen, 1997; Krishnan et al., 1992). A 27 PBPK model intended for use in high-dose to low-dose extrapolation should have the equations 28 and parameters describing dose-dependent phenomena if dose-dependence occurs in the range of 29 interest or assessment. Because the determinants of nonlinear behavior may not be identical 30 across species and age groups (e.g., maximal velocity for metabolism, glutathione 31 concentrations), these parameters are required for each species/age group. During the conduct of 6/29/05 4-9 DRAFT—DO NOT CITE OR QUOTE ------- 1 high-dose to low-dose extrapolation, no change in the parameters of PBPK model is required 2 except for the administered dose or exposure concentration. 3 An example of high-dose to low-dose extrapolation is presented in Figure 4-3. In this 4 figure, both the blood AUC and the amount metabolized over a period of time (12 hr) are plotted 5 as a function of exposure concentrations of toluene. For conducting high-dose to low-dose 6 simulation in this particular example, only the numerical value of the exposure concentration 7 (which is an input parameter for the PBPK model) was changed during every model run. All 8 other model parameters remained the same. The model simulations shown in Figure 4-3 indicate 9 the nonlinear nature of blood AUC as well as the amount of toluene metabolized per unit time in 10 the exposure concentration range simulated. In such cases, the high-dose to low-dose 11 extrapolation of tissue dosimetry should not be conducted assuming linearity; rather it should be 12 performed using tools such as the PBPK models. 13 14 4.3.2. Interspecies Extrapolation 15 The application of PBPK models for interspecies extrapolation of tissue dosimetry is 16 performed in several steps. First, a rodent model is developed to describe the uptake and 17 disposition of the chemical in question by integrating information on the physiological, 18 biochemical, and physicochemical parameters. Then, a priori predictions of the PBPK model 19 are compared with experimental observations to evaluate the adequacy of the structure and the 20 parameter estimates of the rodent model. The next step involves using species-specific or 21 allometrically scaled physiological parameters in the model and replacement of the chemical- 22 specific parameters with appropriate estimates for the species of interest (e.g., humans). Thus, in 23 this approach, the qualitative determinants of pharmacokinetics are considered to be invariant 24 among the various mammalian species. Qualitative differences between species, if any, can also 25 be factored into the existing structure of PBPK models (e.g., if different metabolic pathways 26 existed among species). 27 For conducting interspecies extrapolation of pharmacokinetic behavior of a chemical, 28 quantitative estimates of model parameter values (i.e., partition coefficients, physiological 29 parameters, and metabolic rate constants) in the second species are required. The tissue:air 30 partition coefficients of chemicals appear to be relatively constant across species, whereas 31 blood:air partition coefficients show some species-dependent variability. Therefore, the 32 tissue:blood partition coefficients for human PBPK models have been calculated by dividing the 6/29/05 4-10 DRAFT—DO NOT CITE OR QUOTE ------- 2 3 4 5 6 7 8 9 10 11 12 13 14 15 : i Exposure; ;eonre;ntratrans((;ipm illl-Hlpiltigiljg:;;:. s;;:Exposure j Goncienrt Hatipri ilpgrnji i Figure 4-3. High-dose to low-dose extrapolation of dose metrics using PBPK model for toluene. Inhalation exposures were for 4 hr and AUCs and amount metabolized were calculated for 12 hr. Based on Tardif et al. (1997). rodent tissue:air partition coefficients by the human blood:air partition values (Krishnan and Andersen, 2001). The tissue:air and blood:air partition coefficients for volatile organic chemicals may also be predicted using appropriate data on the content of lipids and water in human tissues and blood (Poulin and Krishnan, 1996a, b). Whereas the adult physiological parameters vary coherently across species, the kinetic constants for metabolizing enzymes do not necessarily follow any type of readily predictable pattern, making the interspecies extrapolation of xenobiotic metabolism difficult. Therefore, the metabolic rate constants for xenobiotics should be obtained in the species of interest. In vivo approaches for determining metabolic rate constants are not always feasible for application in 6/29/05 4-11 DRAFT—DO NOT CITE OR QUOTE ------- 1 humans. The alternative is to obtain such data under in vitro conditions (e.g., Lipscomb et al., 2 1998, 2003). A parallelogram approach may also be used to predict values for the human PBPK 3 model on the basis of metabolic rate constants obtained in vivo in rodents as well as in vitro 4 using rodent and human tissue fractions (Reitz et al., 1988b; Lipscomb et al., 1998). Alternately, 5 for chemicals exhibiting high affinity (low Km) for metabolizing enzymes, Vmax has been scaled 6 to the 0.75 power of body weight, keeping the Km species invariant. This approach may be 7 useful as a crude approximation, but it should be used only when other direct measurements of 8 metabolic parameters are not available or feasible. 9 An example of rat-human extrapolation of the kinetics of toluene using a PBPK model is 10 presented in Figure 2-2. Here the structure of the PBPK model developed in rats was kept 11 unchanged, but the species-specific parameters were determined either by scaling or 12 experimentally, as described above (Tardif et al., 1997). The model was then able to predict 13 accurately the kinetics of toluene in humans (Figure 2-2). Whenever the human data for a 14 particular chemical are not available for evaluation purposes, a corollary approach permitting the 15 use of human data on similar chemicals may be attempted (Jarabek et al., 1994). 16 17 4.3.3. Route-to-Route Extrapolation 18 There are two different approaches to route extrapolation involving PBPK models. The 19 first one is to use an animal model to extrapolate a POD for one route to POD by another route 20 on the basis of equivalent dose metric. The second approach would involve the estimation of the 21 human POD for one route from the available animal POD for another route on the basis of 22 equivalent dose metric. 23 The extrapolation of the kinetic behavior of a chemical from one exposure route to 24 another is performed by adding appropriate equations to represent each exposure pathway. For 25 simulating the intravenous administration of a chemical, a single input representing the dose 26 administered to the animal is included in the equation for mixed venous concentration. Oral 27 gavage of a chemical dissolved in a carrier solvent may be modeled by introducing a first-order 28 or a zero-order uptake rate constant, and dermal absorption has been modeled by including a 29 diffusion-limited compartment to represent skin as a portal of entry (Krishnan and Andersen, 30 2001). After the equations describing the route-specific entry of chemicals into systemic 31 circulation are included in the model, it is possible to conduct extrapolations of pharmacokinetics 32 and dose metrics. This approach is illustrated in Figure 4-4 for inhalation -» oral extrapolation 6/29/05 4-12 DRAFT—DO NOT CITE OR QUOTE ------- Oral c .2 •o 2 JO § g ID > 100 -i 10 1 • 0.1 • 0.01 - 0.001 2345 Time (hours) Inhalation 3 4 5 6 7 8 9 10 11 12 13 i:::::::::;:;;; VeRbus:blftOd:iiS)nc:entfati(jihi: ::;:;;:;::;:;;;::;;:;;;;::;;::::;::(mgA:J ;; 100 - 10 • 1 - 0.1 • r V^ 01234567 Time (hours) Figure 4-4. Oral-to-inhalation extrapolation of the pharmacokinetics of chloroform on the basis of same AUC in blood (7.06 mg/L-hr). The oral dose was 1 mg/kg and the inhaled concentration was 83.4 ppm (4 hr). Based on Corley etal. (1990). of the kinetics of chloroform in rats. For simulating the inhalation pharmacokinetics, the oral dose was set to zero, whereas for simulating chloroform kinetics following oral dosing the inhaled concentration was set to zero (Figure 4-4). Accordingly, 4-hr inhalation exposure to 83.4 ppm chloroform is equal to an oral dose of 1 mg/kg, as determined with PBPK models on the basis of equivalent dose metric (i.e., parent chemical AUC in blood) (Figure 4-4). 6/29/05 4-13 DRAFT—DO NOT CITE OR QUOTE ------- 1 2 3 4 5 6 7 8 9 10 11 43.4. Duration Adjustment On the basis of equivalent dose metric, the duration-adjusted exposure values can be obtained with PBPK models (Andersen et al., 1987; Brodeur et al., 1990). Accordingly, the AUC of a chemical for the exposure duration of the critical study is determined initially using the PBPK model, and then the atmospheric concentration for a continuous exposure (during a day, window of exposure, or lifetime) yielding the same AUC is determined by iterative simulation. Figure 4-5 depicts an example of 4 hr-to-24 hr extrapolation of the pharmacokinetics of toluene in rats, based on equivalent 24-hr AUC (2.4 mg/L-hr). The rats exposed to 50 ppm for 4 hr and 9.7 ppm for 24 hr of toluene would receive the same dose metric. O _l 0 "Si E E 100- 10- 0.1 - 0.01 - 0.001 - 0.0001 RAT 8 12 16 Time (hours) 20 24 HUMAN 12 13 14 15 c o •»= 1 8^! •o g> o E .2 % 8 (U ^> 10 - 1 - 0.1 - 0.01 - 0.001 - n nnm - ^ 0 4 8 12 16 20 2' Time (hours) Figure 4-5. Duration adjustment (4 hr to 24 hr) of toluene exposures in rats, based on equivalent AUC (2.4 mg/L-hr). The rats were exposed to 50 ppm toluene for 4 hr and 9.7 ppm for 24 hr. Based on Tardif et al. (1997). 6/29/05 4-14 DRAFT—DO NOT CITE OR QUOTE ------- 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 4.3.5. Intraspecies Extrapolation Intraspecies extrapolation of the dose metric is conducted using PBPK models with the sole intent of estimating the magnitude of the interindividual variability factor (pharmacokinetic component) for RfC and RfD derivations. In this regard, the population distributions of parameters, particularly those relating to physiology and metabolizing enzymes, are specified in a Monte Carlo approach, such that the PBPK model output corresponds to distributions of dose metric in a population. Using the Monte Carlo approach, repeated computations based on inputs selected at random from statistical distributions for each input parameter (physiological parameters, enzyme content/activity with or without the consideration of polymorphism) are conducted to provide a statistical distribution of the output, i.e., tissue dose. Using the information on the dose metric corresponding to the 95th percentile and 50th percentile for unimodel, normal distribution (Naumann et al., 2001), the magnitude of the inter-individual variability factor can be computed for risk assessment purposes (Figure 4-6). .a OJ .0 o 50th percentile Interindividual variability factor 111 i ' *° Illlli 95th percentile Concentration Figure 4-6. Estimation of the interindividual factor from the 50th (median) and 95th percentile values of a dose metric simulated with a probabilistic PBPK model. 6/29/05 4-15 DRAFT—DO NOT CITE OR QUOTE ------- 1 Even though several past efforts have characterized the impact of the distributions of 2 parameters in adult population, such variability analyses should also account for the life stage- 3 specific changes in physiology, tissue composition, and metabolic activity (reviewed in 4 O'Flaherty, 1994, Corley et al, 2003). 5 6 4.3.6. RfD Derivation 7 When a PBPK model is available for the oral route in test species and adult humans, RfD 8 derivation has been accomplished as follows (e.g., U.S. EPA, 1999b): 9 1. The PBPK model with the parameters of the test species is exercised to determine the 10 dose metric associated with the POD (NOAEL, LOAEL, or BMD), 11 2. The human PBPK model is exercised to determine the oral dose that is associated 12 with the dose metric established in step (1) above, and 13 3. The resulting oral dose is divided by the appropriate uncertainty factors to establish 14 the RfD. 15 16 A variant of the above approach involved dividing the dose metric associated with the 17 POD by the uncertainty factors before exercising the human model to determine equivalent doses 18 (e.g-, Gentry et al., 2002). The outcome is essentially the same if the uncertainty factors are 19 being applied to doses (internal or external) in the linear range of pharmacokinetic processes in 20 the species of interest. Clewell et al., (2002) suggest that the uncertainty factors are more 21 appropriately applied to the dose metric rather than to the external dose. The difference in 22 outcome between the two approaches is likely to exist if the POD is in the partially or completely 23 saturating range. Because the Agency has traditionally applied the uncertainty factors to the 24 external dose and not to the internal dose, it may be useful to undertake a systematic evaluation 25 of the outcome of applying the uncertainty factors to the external and internal doses for various 26 chemicals and situations. 27 In the PBPK model-based approach to RfD derivation, the magnitude of uncertainty 28 factors used is identical to that of the conventional approach except that the animal-to-human 29 uncertainty factor is reduced to three to account for the pharmacodynamic aspect (e.g., Clewell et 30 al., 1997; Gentry et al., 2002). 31 The RfD derivation for ethylene glycol monobutyl ether exemplifies the current approach 32 of PBPK model application (U.S. EPA, 1999b). hi this case, the LOAEL identified in an animal 6/29/05 4-16 DRAFT—DO NOT CITE OR QUOTE ------- 1 study (59 mg/kg-d) was provided as input to the PBPK model to determine the maximal 2 concentration of the metabolite, butoxy acetic acid in blood (BAAmax) (Corley et al., 1997). The 3 dose metric (BAAmax) associated with LOAEL was established in the test species (103 yuM). The 4 human PBPK model was then run to determine the exposure dose that would give the target dose 5 metric (103 |JM) (Corley et al., 1997). The resulting human-equivalent dose of 7.6 mg/kg-d was 6 divided by the appropriate uncertainty factors (30; 10 for interindividual differences and 3 for 7 LOAEL-to-NOAEL extrapolation) to establish the RfD (0.3 mg/kg-d). In this particular case, 8 the interspecies pharmacodynamic factor was set to 1 because in vitro studies suggested that 9 humans are less sensitive than rats to the hematologic effects of ethylene glycol monobutyl ether 10 (U.S. EPA, 1999b). 11 When the BMD is available, a similar approach is used to establish the RfD. In the case 12 of ethylene glycol monobutyl ether, initially the dose metric associated with the BMD was 13 established (BAAmax = 64 /uM) and then the human PBPK model was used to back-calculate the 14 equivalent dose (5.1 mg/kg-d). Using the appropriate uncertainty factor (10 for interindividual 15 variability), the RfD was derived (0.5 mg/kg-d) (U.S. EPA, 1999b). If the human PBPK model 16 was probabilistic in nature, accounting for the population distribution of parameters 17 (biochemical, physiological, and physicochemical), the pharmacokinetic component of the 18 interindividual variability factor could be addressed. Such a PBPK model-based assessment of 19 interindividual variability factor in the derivation of the RfD has been shown with methyl 20 mercury (Clewell et al., 1999). 21 22 4.3.7. RfC Derivation 23 When an inhalation PBPK model is available for the test species and humans, the HEC is 24 derived as follows: 25 1. The PBPK model with the parameters of the test species is run to determine the dose 26 metric associated with the POD (unadjusted or duration-adjusted NOAEL, LOAEL, 27 BMC), and 28 2. The PBPK model with the parameters for an average individual is run to determine 29 the continuous exposure concentration that is associated with the dose metric 30 established in step (1) above. 31 32 Subsequently, appropriate uncertainty factors are used the same way as in the conventional 33 approach (e.g., Clewell et al., 1997; Gentry et al., 2002; also see Section 4.3.6). 6/29/05 4-17 DRAFT—DO NOT CITE OR QUOTE ------- 1 The RfC derivations for m-xylene and vinyl chloride exemplify the application of PBPK 2 models. In the case of m-xylene, using the adjusted NOAEL of 39 mg/m3 as input to the rat 3 PBPK model, the steady-state blood concentration was established (0.144 mg/L) (Tardif et al., 4 1997). The human model was then run to determine the exposure concentration yielding that 5 same dose metric (HEC = 41 mg/m3) (U.S. EPA, 2003). In an alternative approach, the dose 6 metric associated with the unadjusted NOAEL (217 mg/m3, 6 hr/d, 5 d/wk, 13 wks) in the rat 7 was determined using the PBPK model (time-weighted average blood concentration = 0.198 8 mg/L). Then, the human PBPK model was used to determine the 24-hr exposure concentration 9 that would produce this target dose metric (39 mg/m3). Dividing this value by the appropriate 10 uncertainty factors (3 for interspecies pharmacodynamic differences, 10 for interindividual 11 variability, 3 for subchronic to chronic extrapolation, and 3 for database deficiency), the RfC was 12 determined (0.1 mg/m3). 13 In the case of vinyl chloride, the RfC was derived from the NOAEL for the oral route 14 (U.S. EPA, 2000b). The PBPK model was initially used to derive the dose metric associated 15 with the rat NOAEL (0.13 mg/kg-d). Then, the human PBPK model was exercised to determine 16 the continuous exposure concentration associated with the same dose metric (2.5 mg/m3) 17 (Clewell et al., 1995). Using a total uncertainty factor of 30 (3 for toxicodynamic component of 18 IUF, 10 for interindividual variability factor), the RfC was established (0.1 mg/m3). 19 If the available human PBPK model is probabilistic in nature, accounting for the 20 population distribution of parameters (biochemical, physiological, and physicochemical), the 21 magnitude of the interindividual variability factor can be estimated (Delic et al., 2000). In that 22 case, the interindividual variability factor will be set to 3 (to account only for pharmacodynamic 23 differences). 24 25 4.3.8. Cancer Risk Assessment (unit risk estimates, RfC, and RfD) 26 When data for the MOA in a cancer risk assessment suggest a threshold (curvilinear) 27 dose-response relationship, the applications of PBPK models are similar to those in the 28 reference dose process (i.e., RfD and RfC derivation). If a linear dose-response model is 29 more applicable, or the MOA is unknown, then the following steps describe uses for a PBPK 30 model (e.g., Andersen et al., 1987): 31 32 6/29/05 4-18 DRAFT—DO NOT CITE OR QUOTE ------- 2 1. The relationship between administered dose or exposure concentration and dose 3 metric is established for the test species using PBPK model, 4 2. The quantitative relationship between dose metric and the cancer incidence observed 5 in the bioassay(s) is characterized to estimate the dose metric-based slope factor, and 6 3. The potential doses or exposure concentrations yielding the dose metric associated 7 with various levels of risk (e.g., 1 x 10~6) are back-calculated using the human PBPK 8 model. 9 10 For assessing the cancer risk associated with human exposures, the exposure 11 concentration is provided as input to human PBPK model to simulate the dose metric, which is 12 then multiplied with the dose metric-based slope factor. In the cancer risk assessments using 13 PBPK models, it is assumed that the tissue response associated with a given level of dose metric 14 in the target tissue is the same in test animals and in humans (e.g., Andersen et al., 1987). It is a 15 reasonable assumption that can be revised as a function of species-specific mechanistic 16 information available for a given chemical. 17 The demonstration of the applicability of PBPK models in cancer risk assessment was 18 first accomplished with dichloromethane, which caused liver and lung tumors in mice exposed to 19 2,000 or 4,000 ppm 6 hr/d, 5 d/wk for lifetime (Andersen et al., 1987). In this case, the mouse 20 PBPK model was used to calculate the tissue dose of metabolites and parent chemical arising 21 from exposure scenarios comparable to those of the cancer bioassay study, and their relationship 22 to the observed tumor incidence was then examined. Because the parent chemical is nonreactive, 23 Andersen et al. (1987) considered it an unlikely candidate responsible for the tumorigenicity. 24 Hence the relationship between the tissue exposure to its metabolites and tumor incidence was 25 examined (Table 4-2). Whereas the dose metric based on oxidative pathway varied little 26 between 2,000 and 4,000 ppm, the flux through the glutathione pathway increased with 27 increasing dose of dichloromethane and corresponded well with the degree of dichloromethane- 28 induced tumors at these exposure concentrations. 29 The model prediction of the target tissue dose of the glutathione conjugate resulting from 30 6-hour inhalation exposures to 1-4,000 ppm dichloromethane is presented in Figure 4-7 31 (Andersen etal., 1987). The estimation of target tissue dose of dichloromethane-glutathione 32 conjugate by linear back-extrapolation gives rise to a 21-fold higher estimate than that obtained 33 by the PBPK modeling approach. This discrepancy arises from the nonlinear behavior of 6/29/05 4-19 DRAFT—DO NOT CITE OR QUOTE ------- 1 2 3 Table 4-2. Relationship between tumor prevalence and dichloromethane metabolites produced by microsomal and glutathione pathway for the bioassay conditions (methylene chloride-dose response in female mice) Exposure (ppm) 0 2,000 4,000 Microsomal pathway dose8 Liver — 3,575 3,701 (Lung) — (1,531) (1,583) Gluthathione pathway dose8 Liver — 851 1,811 (Lung) — (123) (256) Tumor prevalence Liver 6 33 83 (Lung) (60) (63) (85) 4 5 6 7 a Tissue dose is cumulative daily exposure (mg metabolized/volume tissue/day). Source: Adapted from Andersen et al. (1987). 8 9 10 11 12 13 14 15 16 17 innn - to -inn . H Pathway (mg/L/d 0 I 0.1 • 0.01 • /X X X /..-'•- / X XX ** Y />' ^ • •* X S 10 100 1000 10C PPM in air Figure 4-7. PBPK model predictions of glutathione-pathway metabolites in liver in mice. The three curves are for a linear extrapolation from the bioassay exposures of 2,000 and 4,000 ppm (upper curve), the expected tissue dose-based on model parameters for the mouse (middle line), and the expected dose expected in humans, based on human model parameters (bottom line). Similar curves were published in Andersen et al. (1987). The curvature occurs because oxidation reactions that are favored at low inhaled concentrations become saturated as inhaled concentration increases above several hundred ppm. 6/29/05 4-20 DRAFT—DO NOT CITE OR QUOTE ------- 1 dichloromethane metabolism at high-exposure concentrations. At exposure concentrations 2 exceeding 300 ppm, the cytochrome P-450-mediated oxidation pathway is saturated, giving rise 3 to a corresponding disproportionate increase in the flux through glutathione conjugation 4 pathway. By accounting for the species-specific differences in metabolism rates and physiology 5 in the PBPK model, the target tissue dose for humans was estimated to be some 2.7 times lower 6 than that for the mouse. The target tissue dose-based slope factor has subsequently been used for 7 characterizing the cancer risk associated with human exposures (Andersen et al., 1987; Reitz et 8 al., 1989; Haddad et al., 2001). The case of dichloromethane exemplifies how PBPK models can 9 be used for improving the dose-response relationship on the basis of appropriate dose metrics, 0 thus leading to scientifically sound conduct of interspecies and high-dose to low-dose 11 extrapolations essential for cancer risk assessments. 12 13 4.3.9. Mixture Risk Assessment 14 PBPK. models permit the risk assessment of chemical mixtures by facilitating the 15 simulation of change in dose metrics due to muitichemical interactions (Haddad et al., 2001). 16 For conducting tissue dosimetry-based assessments for mixtures, adequately evaluated PBPK 17 models for the mixture in the test species and in humans are required. Further, the health- 18 protective values for the individual chemicals (e.g., slope factor, RfD, RfC) should be known. 19 The approach for using PBPK models in risk assessment of mixtures of systemic toxicants or 20 carcinogens exhibiting threshold mechanism of action, would consist of (Haddad et al., 2001): 21 1. Characterizing the dose metrics associated with the RfC or RfD of mixture 22 components, 23 2. Obtaining predictions of dose metrics of each mixture component in humans, based 24 on information on exposure levels provided as input to the mixture PBPK model; and 25 3. Determining the sum total of the ratios of the results of steps (1) and (2) for each 26 component during mixed exposures. 27 28 Notationally, for chemicals in mixture acting on the same target organ: " C TJJ ^~* tissue,exp,i. fll — / • C '=1 tissue, ref ,i. 29 30 6/29/05 4-21 DRAFT—DO NOT CITE OR QUOTE ------- 1 2 where HI is the hazard index, CtisSUe;eXp,i refers to the tissue concentration of the dose metric of 3 component / predicted in the experimental animal at the POD using PBPK model that accounts 4 for multiple interactions occurring in the mixture, and Ctissue,ref,/ is the dose metric associated with 5 human exposure to the RfD or RfC of component /'. 6 7 Similarly, for carcinogens with slope factor (Haddad et al., 2001): 8 1. The dose metric-based slope factor should be established for each component using 9 the animal PBPK model, 10 2. The dose metric associated with human exposure concentrations should be 11 established using mixture PBPK models, and 12 3. The results of steps (1) and (2) should be combined to determine the potentially 13 altered cancer response during mixed exposures. 14 15 Notationally, 16 17 18 19 where P(d) is the probability of excess cancer incidence in an exposed population, q*i,tissme = dose 20 metric-based cancer slope factor for each of the mixture components (i = 1,2, ..., n), and d^ssue 21 refers to the dose metric for each mixture component simulated using PBPK models that account 22 for multi-chemical interactions. 23 Risk assessments based on the use of PBPK models for single chemicals and mixtures, as 24 detailed in Section 4.3, account for only the pharmacokinetic aspect or, more specifically, target 25 tissue exposure to toxic moiety. If these tissue exposure simulations are combined with 26 pharmacodynamic models, then better characterization of dose-response relationships and 27 prediction of PODs (NOAEL, BMD, BMC) may become possible. 28 29 4.4. LINKAGE TO PHARMACODYNAMIC MODELS 30 The identification of PODs by simulation may become possible with the use of BBDR 31 models. These models would require the linkage of quantitative descriptions of 32 pharmacokinetics and pharmacodynamics via mechanism of action. Accordingly, the output of 6/29/05 4-22 DRAFT—DO NOT CITE OR QUOTE ------- 1 PBPK models is linked to the pharmacodynamic model using an equation that reflects the 2 researchers' hypothesis of how the toxic chemical participates in the initiation of cellular 3 changes leading to measurable toxic responses. For example, certain DNA adducts cause 4 mutations, cytotoxic metabolites kill individual cells, and expression of growth factors can act as 5 a direct proliferation stimulus. In each of these cases, the temporal change in the dose metric 6 simulated by the PBPK model is linked with mathematical descriptions of the process of adduct 7 formation, cytotoxicity, or proliferation in the BBDR models to simulate the quantitative 8 influence of these processes on tumor outcome. 9 Figure 4-8 presents an example of the relationship between dose metric (simulated by the 10 PBPK model) and fraction of liver cells killed (simulated by pharmacodynamic model) for 11 chloroform. In this case, the pharmacodynamic model consisted of difference equations to 12 simulate time-dependent changes in the number of hepatocytes in the liver as a function of basal 13 14 15 16 17 18 0.90 0.00 1.09 2.00 3,0Q 4,60 5,00 9.90 Liver dose surrogate (|X«K>| metabolized/g liver/60 tniii) Figure 4-8. Relationship between the dose metric (pmol metabolized/g liver/hr) simulated by PBPK model and the cell killing inferred from pharmacodynamic model for chloroform. Source: Page et al. (1997). 6/29/05 4-23 DRAFT—DO NOT CITE OR QUOTE ------- 1 rates of cell division and death, chloroform-induced cytolethality, and regenerative replications 2 (Conolly and Butterworth, 1995; Page et al., 1997). 3 Table 4-3 presents a list of pharmacodynamic models for cancer and noncancer 4 endpoints. A characteristic of several of these pharmacodynamic models is that they are able to 5 simulate the normal physiological processes (e.g., cell proliferation rates, hormonal cycle) and 6 additionally account for the ways in which chemicals perturbate such phenomena, leading to the 7 onset and progression of injury. The pharmacodynamic models, for linkage with PBPK models, 8 are not available for a number of toxic effects and modes of action. This situation is a result, in 9 part, of the complex nature of these models and the extensive data requirements for development 10 and evaluation of these models for various exposure and physiological conditions. 11 With the availability of integrated pharmacokinetic-pharmacodynamic models, the 12 scientific basis of the process of estimating PODs and characterizing the dose-response curve 13 will be significantly enhanced. Additionally, such a modeling framework will facilitate a 14 quantitative analysis of the impact of pharmacodynamic determinants on the toxicity outcome, 15 such that the magnitude of the pharmacodynamic component of the interspecies and intraspecies 16 factors can be characterized more confidently. Even though some PBPK models have been used 17 in RfD, RfC, and unit risk estimate derivation for a number of substances (Table 4-1), the need 18 for applying such models (where possible) should be continuously explored. 19 20 6/29/05 4-24 DRAFT—DO NOT CITE OR QUOTE ------- 1 2 Table 4-3. Examples of biologically based models of endpoints and processes of toxicological relevance Toxicity end point or process Cancer Cholinesterase inhibition Developmental toxicity Estrus cycle Gene expression Granulopoiesis Nephrotoxicity Teratogenic effect Features Simulation of relative roles of initiation, promotion, cyclothality, and proliferation Simulation of dose-dependent inhibition of plasma cholinesterase, red blood cell acetyl cholinesterase and brain acetyl cholinesterase, and nontarget B-esterase Simulation of altered cell kinetics as the biological basis of developmental toxicity Simulation of interactions of estradiol and lutenizing hormone Simulation of induction of CYP1 Al/2 protein expression in multiple tissues Simulation of loss of proliferating cells and loss of functional cells Simulation of induction of renal 2\i. globulin in male rat kidney as a function of proteolytic degradation and hepatic production Sensitivity distribution of embryo as a function of age and stage of development Chemical studied 2-acetylamino fluorine Chloroform Dimethylnitrosamine Formaldehyde PCBs Pentachlorobenzene Saccharin Organophophates Methyl mercury Endocrine-modulating substances TCDD Cyclophosphamide 2,2,4-trimethyl-2-phenol Hydroxyurea References Armitage and Doll (1957); Moolgavkar and Venzon (1 979); Moolgavkar and Knudson (1981); Cohen and Ellwein (1990); Moolgavkar & Luebeck (1 990); Luebeck et al. (1991); Chen (1993); Conolly and Andersen (1997); Conolly and Kimbell (1994); Thomas et al. (2000); Conolly et al. (2003); Tan et al. (2003) Gearhart et al. (1990, 1994); Timchalk et al. (2002) Leroux et al. (1996); Faustman etal. (1999) Andersen et al. (1997) Santostefano et al. (1998) Steinbach et al. (1980) Kohn and Melnick (1 999) Luecke etal. (1997) 3 4 5 6/29/05 4-25 DRAFT—DO NOT CITE OR QUOTE ------- GLOSSARY Absorbed dose: The amount crossing a specific absorption barrier (e.g., the exchange boundaries of the skin, lung, and digestive tract) through uptake processes. Applied dose: The amount presented to an absorption barrier and available for absorption (although not necessarily having yet crossed the outer boundary of the organism). Area under the curve: The concentration of a chemical in tissues or blood integrated over time. It is a measure of tissue exposure to chemicals over a period of time. Bayesian statistics: An approach that considers a model's parameters as random variables with a probability distribution for describing each parameter. The distribution based only on prior information and assumptions is called the prior distribution. Analysis of new data yields a posterior distribution that reconciles the prior information and assumptions with the new data. Benchmark dose (BMD) or benchmark concentration (BMC): A dose or concentration that produces a predetermined change in response rate of an adverse effect (called the benchmark response) compared to background. Biologically based dose response model: A predictive model that describes biological processes at the cellular and molecular level linking the target organ dose to the adverse effect. Cancer scope factor: An estimate of the increased cancer risk from a lifetime exposure to an agent. This estimate, usually expressed in units of proportion (of a population) affected, is generally reserved for use in the low-dose region of the dose-response relationship. It is often an upper bound, approximating a 95% confidence limit. Clearance: Volume containing the amount of drug eliminated per unit time by a specified organ; it has the dimension of a flow per unit time. Critical effect: The first adverse effect, or its known precursor, that occurs to the most sensitive species as the dose rate of an agent increases. Delivered dose: The amount of a substance available for interactions with biologically significant receptors in the target organ. Dose metric: The target tissue dose that is closely related to ensuing adverse responses. Dose metrics, used for risk assessment applications, should reflect the biologically active form of chemical, its level, and duration of internal exposure, as well as intensity. Dose-response assessment: The process of determining the relationship between the magnitude of administered, applied, or internal doses and biological responses. Response can be expressed as measured or observed incidence or change in level of response, percent response in groups of subjects (or populations), or the probability of occurrence or change in level of response within a population. 6/29/05 G-l DRAFT—DO NOT CITE OR QUOTE ------- Exposure assessment: The process of identifying and evaluating the human population exposed to a toxic agent by describing its composition and size, as well as the type, magnitude, frequency, route, and duration of exposure. Half-life: Interval of time required for one-half of a given substance present in an organ to leave it through processes other than physical decay. It is a constant only for mono-exponential functions. Human equivalent concentration (HEC): The human concentration (for inhalation exposure) of an agent that is believed to induce the same magnitude of toxic effect as the exposure concentration in experimental animal species. This adjustment may incorporate pharmacokinetic information on the particular agent, if available, or use a default procedure. Internal dose: A more general term denoting the amount absorbed without respect to specific absorption barriers or exchange boundaries. The amount of the chemical available for interaction by any particular organ or cell is termed the delivered or biologically effective dose for that organ or cell. Markov-chain Monte-Carlo simulation: An approach that has frequently been used within a Bayesian statistical framework to (i) sample each model's parameters from their prior distributions, (ii) fit the model with the sampled parameters to several additional experimental data sets, and (iii) compare the model's predictions with the experimental results to obtain posterior distributions for the model's parameters that improve the model's fit. These steps are repeated thousands of times until each parameter's posterior distribution converges to a more robust distribution that reflects a wider database. Pharmacokinetic models: Mathematical descriptions simulating the relationship between external exposure levels and the biologically effective dose at a target tissue. Pharmacokinetic models take into account absorption, distribution, metabolism, and elimination of the administered chemical and its metabolites. Pharmacodynamic models: Mathematical descriptions simulating the relationship between a biologically effective dose and the occurrence of a tissue response. Physiologically based pharmacokinetic (PBPK) model: A model that estimates the dose to target tissue by taking into account the rate of absorption into the body, distribution and storage in tissues, metabolism, and excretion on the basis of interplay among critical physiological, physicochemical, and biochemical determinants. Point of departure: The dose-response point that marks the beginning of a low-dose extrapolation. This point can be the lower bound on dose for an estimated incidence or a change in response level from a dose-response model (BMD, BMC), or a NOAEL or LOAEL for an observed incidence or change in level or response. Potential dose: The amount ingested, inhaled, or applied to the skin. Reference concentration (RfC): An estimate (with uncertainty spanning perhaps an order of magnitude) of a continuous inhalation exposure to the human population (including sensitive 6/29/05 G-2 DRAFT—DO NOT CITE OR QUOTE ------- subgroups) that is likely to be without an appreciable risk of deleterious effects during a lifetime. It can be derived from a NOAEL, LOAEL, or benchmark concentration, with uncertainty factors generally applied to reflect limitations of the data used. Reference dose (RfD): An estimate (with uncertainty spanning perhaps an order of magnitude) of a daily oral exposure to the human population (including sensitive subgroups) that is likely to be without an appreciable risk of deleterious effects during a lifetime. It can be derived from a NOAEL, LOAEL, or benchmark dose, with uncertainty factors generally applied to reflect limitations of the data used. Steady state: A variable is said to have attained steady state when its value stays constant in a given interval of time, i.e., when its derivative is zero. Target organ: The biological organ(s) most adversely affected by exposure to a chemical or physical agent Terminal half-life: The terminal half-life is the interval of time for the concentration of the drug in a compartment to decrease 50% in its final phase. Uncertainty: Uncertainty occurs because of lack of knowledge. Uncertainty can often be reduced with greater knowledge of the system or by collecting more and better experimental or simulation data. Uncertainty/variability factors: Generally 10-fold, default factors used in operationally deriving the RfD and RfC from experimental data. The factors are intended to account for (i) variation in sensitivity among the members of the human population (i.e., interindividual variability), (ii) uncertainty in extrapolating animal data to humans (i.e., interspecies uncertainty), (iii) uncertainty in extrapolating from data obtained in a study with less-than- lifetime exposure to lifetime exposure (i.e., extrapolating from subchronic to chronic exposure), (iv) uncertainty in extrapolating from a LOAEL rather than from a NOAEL, and (v) uncertainty associated with extrapolation when the database is incomplete. Variability: Variability refers to true heterogeneity or diversity. Differences among individuals in a population are referred to as interindividual variability; differences for one individual over time are referred to as intraindividual variability. Volume of distribution: The volume of distribution is the ratio between the administered dose and plasma or blood concentration of a chemical. 6/29/05 G-3 DRAFT—DO NOT CITE OR QUOTE ------- REFERENCES Andersen, ME. (1995) Development of physiologically based pharmacokinetic and physiologically based pharmacodynamic models for applications in toxicology and risk assessment. Toxicol Lett 79:35^44. Andersen, ME; Dennison, JE. (2001) Mode of action and tissue dosimetry in current and future risk assessments. Sci Tot Environ 274:3-14. Andersen, ME and Jarabek, AM. (2001) Nasal tissue dosimetry-issues and approaches for "Category!" gases: a report on a meeting held in Research Triangle Park, NC, February 11-12,1998. Inhal Toxicol 13:415-435. Andersen, ME; Gargas, ML; Jones, RA; et al. (1980) Determination of the kinetic constants for metabolism of inhaled toxicants in vivo by gas uptake measurements. Toxicol Appl Pharmacol 54:100-116. Andersen, ME; Clewell, HJ, III; Gargas, ML; et al. (1987) Physiologically-based pharmacokinetics and risk assessment process for methylene chloride. Toxicol Appl Pharmacol 87:185-205. Andersen, ME; Clewell, HJ, III; Gargas, ML. (1991) Physiologically-based pharmacokinetic modeling with dichloromethane, its metabolite carbon monoxide and blood carboxyhemoglobin in rats and humans. Toxicol Appl Pharmacol 108:14-27. Andersen, ME; Mills, JJ; Gargas, ML; et al. (1993) Modeling receptor-mediated processes with dioxin: implications for pharmacokinetics and risk assessment. Risk Anal 13:25-3 6. Andersen, ME; Clewell, HJ, III; Frederick, CB. (1995) Applying simulation modeling to problems in toxicology and risk assessment-a short perspective. Toxicol Appl Pharmacol 133:181-187. Andersen, ME; Clewell, HJ, III; Gearhart, J; et al. (1997) Pharmacodynamic model of the rat estrus cycle in relation to endocrine disrupters. J Toxicol Environ Health 52:189-209. Andersen, ME; Sarangapani, R; Gentry, PR; et al. (1999) Application of a hybrid CFD-PBPK nasal dosimetry model in an inhalation risk assessment: an example with acrylic acid. Toxicol Sci 57:312-325. Andersen, ME; Green, T; Frederick, CB; et al. (2002) Physiologically based pharmacokinetic (PBPK) models for nasal tissue dosimetry of organic esters: assessing the state-of-knowledge and risk assessment applications with methyl methacrylate and vinyl acetate. Regul Toxicol Pharmacol 36:234-245. Armitage, P; Doll, R. (1957) A two-stage theory of carcinogenesis in relation to the age distribution of human cancer. Brit J Cancer 11:161-169. Arms, AD; Travis, CC. (1988) Reference physiological parameters in pharmacokinetic modeling. U.S. Environmental Protection Agency, Office of Health and Environmental Assessment, Washington, DC. EPA/600/6- 88/004. Asgharian, B; Wood, R; Schlesinger, RB. (1995) Empirical modeling of particle deposition in the alveolar region of the lungs: a basis for interspecies extrapolation. Fund Appl Toxicol 27:232-238. Ashani, Y; Pistinner, S. (2004) Estimation of the upper limit of human butylcholinesterase dose required for protection against organophosphates toxicity: a mathematically based toxicokinetic model. Toxicol Sci 77:358-367. Aylward, LL; Hays, SM; Karch, NJ; et al. (1996) Relative susceptibility of animals and humans to the cancer hazard posed by 2,3,7,8-tetrachlorodibenzo-p-dioxin using internal measures of dose. Environ Sci Technol 30:3534-3543. Ball, R; Schwartz, SL. (1994) Cmatrix: software for physiologically based pharmacokinetic modeling using a symbolic matrix representation system. Comput Biol Med 24:269-276. 6/29/05 R-l DRAFT—DO NOT CITE OR QUOTE ------- Barnes, DG; Dourson, M. (1988) Reference dose (RfD): description and use in health risk assessments. Regul Toxicol Pharmacol 8:471-495. Barton, HA; Clewell, HJ, III. (2000) Evaluating noncancer effects of trichloroethylene: dosimetry, mode of action, and risk assessment. Environ Health Perspect 108(2):323-334. Beliveau, M; Krishnan, K. (2000) Estimation of rat blood: air partition coefficients of volatile organic chemicals using reconstituted mixtures of blood components. Toxicol Lett 116:183-188. Beliveau, M; Tardif, R; Krishnan, K. (2003) Quantitative structure-property relationships for physiologically based pharmacokinetic modeling of volatile organic chemicals in rats. J Toxicol Appl Pharmacol 189:221-232. Benignus, VA; Boyes, WK; Bushnell, PJ. (1998) A dosimetric analysis of behavioral effects of acute toluene exposure in rats and humans. Toxicol Sci 43:186-195. Bemillon, P; Bois, FY. (2000) Statistical issues in toxicokinetic modeling: a Bayesian perspective. Environ Health Perspect (Suppl 108):883-893. Blancato, JN; Saleh, MA. (1994) Physiologically based pharmacokinetic models. Examples of their use in exposure and risk assessment. Saleh, MA; Blancato, JN; Nauman, CH. eds. In: Biomarkers of human exposure to pesticides. Washington, DC: American Chemical Society; pp. 264-283. Bogaards, JJ; Freidig, AP; van Bladeren, PJ. (2001) Prediction of isoprene diepoxide levels in vivo in mouse, rat and man using enzyme kinetic data in vitro and physiologically-based pharmacokinetic modelling. Chem Biol Interact 138:247-265. Bogdanffy, MS; Sarangapani, R. (2003) Physiologically-based kinetic modeling of vapours toxic to the respiratory tract. Toxicol Lett 138:103-117. Bogdanffy, MS; Sarangapani, R; Plowchalk, DR; et al. (1999) A biologically risk assessment for vinyl acetate- induced cancer and noncancer inhalation toxicity. Toxicol Sci 51:19-35. Bogdanffy, MS; Plowchalk, DR; Sarangapani, R; et al. (2001) Mode-of-action-based dosimeters for interspecies extrapolation on vinyl acetate inhalation risk. Inhal Toxicol 13:377-396. Bois, FY. (1999) Analysis of PBPK models for risk characterization. AnnN Y Acad Sci 895:317-337. Bois, FY. (2000a) Statistical analysis of Fisher et al. PBPK model of trichloroethylene kinetics. Environ Health Perspect 108(Suppl 2):275-282. Bois, FY. (2000b) Statistical analysis of Clewell et al. PBPK model of trichloroethylene kinetics. Environ Health Perspect 108(Suppl 2):307-316. Bois, FY; Woodruff, TJ; Spear, RC. (1991) Comparison of three physiologically based pharmacokinetic models for benzene disposition. Toxicol Appl Pharmacol 110:79-88. Bouchard, M; Brunet, RC; Droz, PO; et al. (2001) A biologically based dynamic model for predicting the disposition of methanoland its metabolites in animals and humans. Toxicol Sci 64:169-184. Boyes, WK; Bussnell, PJ; Crofton, KM; et al. (2000) Neurotoxic and pharmacokinetic responses to trichloroethylene as a function of exposure scenario. J Toxicol Environ Health 108:317-322. Brodeur, J; Lapare', S; Krishnan, K; et al. (1990) Le probleme de 1'ajustement des valeurs limites d'exposition pour des horaires de travail non-conventionnels: utilit6 de la modelisation pharmacocinetique a base physiologique. Travail et Sante 6(2):SI 1-16. 6/29/05 R-2 DRAFT—DO NOT CITE OR QUOTE ------- Brown, RP; Delp, MD; Lindstedt, SL; et al. (1997) Physiological parameter values for physiologically based pharmacokinetic models. Toxicol Ind Health 13:407-484. Burmaster, DE; Murray DMA. (1998) Trivariate distribution for the height, weight, and fat of adult men. Risk Anal 8:385-389. Bush, ML; Frederick, CB; Kimbell, JS; et al. (1998) A CFD-PBPK hybrid model for simulating gas and vapor uptake in the rat nose. Toxicol Appl Pharmacol 150:133-145. Bushnell, PJ. (1997) Concentration time relationships for the effects of inhaled trichloroethylene on signal detection behavior in the rats. Fundam Appl Toxicol 36:30-38. Canuel, G; Viau, C; Krishnan, K. (2000) A modeling framework for back-calculating ambient concentrations from data on biomarkers: proceedings of the International Conference on Health Sciences Simulation; January 27-29; San Diego, CA. The Society for Computer Simulation International. Casanova, M; Conolly, RB; Heck, HA. (1996) DNA-protein cross-links (DPX) and cell proliferation in B6C3F, mice but not Syrian golden hamsters exposed to dichloromethane: pharmacokinetics and risk assessment with DPX as dosimeter. Fund Appl Toxicol 31:103-116. Chen, CW. (1993) Armitage-Doll two-stage model: implications and extension. Risk Anal 13:273-279. Chen, HSG; Gross, JF. (1979) Estimation of tissue to plasma partition coefficients used in physiological pharmacokinetic models. JPharmacokinetBiopharm7:117-125. Clarke, DO; Duignan, JM; Welsch, F. (1992) 2-Methoxyacetic acid dosimetry-teratogenicity relationships in CD-I mice exposed to 2-methoxyethanol. Toxicol Appl Pharmacol 114:77. Clarke, DO; Elswick, BA; Welsch, F; et al. (1993) Pharmacokinetics of 2-methoxyethanol and 2-methoxyacetic acid in the pregnant mouse: a physiologically-based mathematical model. Toxicol Appl Pharmacol 121:239-252. Clark, LH; Setzer, RW; Barton, HA. (2004) Framework for evaluation of physiologically-based pharmacokinetic models for use in safety or risk assessment. Risk Anal 24(6)1697—1717. Clewell, HJ, III; Andersen, ME. (1987) Dose, species and route extrapolation using physiologically-based pharmacokinetic models. Drink Water and Health 8:159-182. Clewell, HJ, III; Jarnot, BM. (1994) Incorporation of pharmacokinetics in noncancer risk assessment: example with chloropentafluorobenzene. Risk Anal 14:265-276. Clewell, HJ, III; Lee, TS; Carpenter, RL. (1994) Sensitivity of physiologically based pharmacokinetic models to variation in model parameters—methylene chloride. Risk Anal 14:521-531. Clewell HJ; Gentry PR; Gearhart JM; et al. (1995) Considering pharmacokinetic and mechanistic information in cancer risk assessments for environmental contaminants: examples with vinyl chloride and trichloroethylene. Chemosphere 31:2561-2578. Clewell, HJ, III; Gentry, PR; Covington, TR; et al. (2000) Development of a physiologically based pharmacokinetic model of trichloroethylene and its metabolites for use in risk assessment. Environ Health Perspect 108:283-305. Clewell, HJ, III; Gentry, PR; Gearhart, JM; et al. (2001) Comparison of cancer risk estimates for vinyl chloride using animal and human data with a PBPK model. Sci Total Environ 274:37-66. Clewell, HJ, III; Andersen, ME; Barton, HA. (2002) A consistent approach for the application of pharmacokinetic modeling in cancer and noncancer risk assessment. Environ Health Perspect 110:85-93. 6/29/05 R-3 DRAFT—DO NOT CITE OR QUOTE ------- Clewell, HJ, III; Teeguarden, JG; McDonald, T; et aL (2002) Review and evaluation of the potential impact of age- and gender-specific pharmacokinetic differences on tissue dosimetry. CritRev Toxicol 32:329-389. Cohen, SM; Ellwein, LB. (1990) Cell proliferation in carcinogenesis. Science 249:1007-1011. Cole, CE; Tran, HT; Schlosser, PM. (2001) Physiologically based pharmacokinetic modeling of benzene metabolism in mice through extrapolation from in vitro to in vivo. J Toxicol Environ Health 62:439-465. Collins, AS; Sumner, SC; Borghoff, SJ; Medinsky, MA. (1999) A physiological model for tert-amyl methyl ether and tert-amyl alcohol: hypothesis testing of model structures. Toxicol Sci 49:15-28. Collins, J. (1987) Prospective predictions and validations in anti-cancer therapy. In: Pharmacokinetics in risk assessment drinking water and health. Washington, DC: National Academy Press. Conolly, RB: Andersen, ME. (1997) Hepatic foci in rats after diethyl-nitrosamine initiation and 2,3,7,8- tetrachlorodibenzo-p-dioxin promotion: evaluation of a quantitative two-cell model and of C YP 1 Al/1 A2 as a dosimeter. Toxicol Appl Pharmacol 146:281-293. Conolly, RB; Butterworth, BE. (1995) Biologically based dose response model for hepatic toxicity: a mechanistically based replacement for traditional estimates of noncancer risk. Toxicol Lett (82/83):901—906. Conolly, RB; Kimbell, JS. (1994) Computer simulation of cell growth governed by stochastic processes: application to clonal growth cancer models. Toxicol Appl Pharmacol 124:284-295. Conolly, RB; Limbell, JS; Janszen, D; et al. (2003) Biologically motivated computational modeling of formaldehyde carcinogenicity in the F344 Rat. Toxicol Sci 75:432-447. Corley, RA; Mandrela, AL; Smith, FA. (1990) Development of a physiologically based pharmacokinetic model for chloroform. Toxicol Appl Pharmacol 103:512-527. Corley, RA; Markham, DA; Banks, C; et al. (1997) Physiologically based pharmacokinetics and the dermal absorption of 2-butoxyethanol vapor by humans. Fund Appl Toxicol 39:120-130. Corley, RA; Gordon, SM; Wallace, LA. (2000) Physiologically based pharmacokinetic modeling of the temperature- dependent dermal absorption of chloroform by humans following bath water exposures. Toxicol Sci 53:13-23. Corley, RA; Mast, TJ; Carney, EW; et al. (2003) Evaluation of physiologically based models of pregnancy and lactation for their application in children's health risk assessments. Crit Rev Toxicol 33:137-211. Cox, LA, Jr. (1996) Reassessing benzene risks using internal doses and Monte Carlo uncertainty analysis. Environ Health Perspect 104:1413-1429. Cruzen, G; Carlson, GP; Johnson, KA; et al. (2002) Styrene respiratory tract toxicity and mouse lung tumors are mediated by CYP2F-generated metabolites. Reg Toxicol Pharmacol 35:308-319. Csanady, GA; Kreuzer, PE; Baur, C; et al. (1996) A physiological toxicokinetic model for 1,3-butadiene in rodents and man: blood concentrations of 1,3-butadiene, its metabolically formed epoxides, and of haemoglobin adducts- relevance of glutathione depletion. Toxicology 113:300-305. Dankovic, DA; Bailer, AJ. (1994) The impact of exercise and intersubject variability on dose estimates for dichloromethane derived from a physiologically based pharmacokinetic model. Fund Appl Toxicol 22:20-25. Davies, B; Morris, T. (1993) Physiological parameters in laboratory animals and humans. Pharm Res 10:1093-1095. De Jongh, J; Blaauboer, BJ. (1996) Simulation of toluene kinetics in the rat by a physiologically based pharmacokinetic model with application of biotransformation parameters derived independently in vitro and in vivo. Fund Appl Toxicol 32:260-268. 6/29/05 R-4 DRAFT—DO NOT CITE OR QUOTE ------- De Jongh, J; Blaauboer, BJ. (1997) Simulation of lindane kinetics in rats. Toxicology 122:1-9. Delic, JI; Lilly, JD; MacDonald, AJ; et al. (2000) The utility of PBPK in the safety assessment of chloroform and carbon tetrachloride. Reg Toxicol Pharmacol 32:144-155. Dong, MH. (1994) Microcomputer programs for physiologically-based pharmacokinetic (PB-PK) modelling. Comput Methods Programs Biomed 45:213-221. Dome, JL; Walton, K; Renwick, AG. (2001 a) Uncertainty factors for chemical risk assessment: human variability in the pharmacokinetics of CYP1A2 probe substrates. Food Chem Toxicol 39:681-696. Dome, JL; Walton, K; Renwick, AG. (2001 b) Human variability in glucuronidation in relation to uncertainty factors for risk assessment. Food Chem Toxicol 39:1153-1173. Dome, JL; Walton, K; Slob, W; et al. (2002) Human variability in polymorphic C YP2D6 metabolism: is the kinetic default uncertainty factor adequate? Food Chem Toxicol 40:1633-1656. Dourson, ML; Knauf, LA; Swartout, JD. (1992) On reference (RfD) and its underlying toxicity data base. Toxicol Ind Health 8:171-189. Droz, PO; Berode, M; Jang, JY. (1999) Biological monitoring of tetrahydrofuran: contribution of a physiologically based pharmacokinetic model. Am Ind Hyg Assoc J 60:243-248. Easterling, MR; Evans, MV; Kenyon, EM. (2000) Comparative analysis of software for physiologically based pharmacokinetic modeling: simulation, optimization, and sensitivity analysis. Toxicol Methods 10:203-229. Edler, L. (1999) Uncertainty in biomonitoring and kinetic modeling. AnnN Y Acad Sci 895:80-100. El-Masri, HA; Bell, DA; Portier, CJ. (1999) Effects of glutathione transferase theta polymorphism on the risk estimates of dichloromethane to humans. Toxicol Appl Pharmacol 158:221-230. Farrar, D; Allen, B; Crump, K; et al. (1989) Evaluation of uncertainty in input parameters to pharmacokinetic models and the resulting uncertainty in output. Toxicol Let 49:371-385. Farris, FF; Dedrick, RL; King, FG. (1988) Cisplatin pharmacokinetics: applications of a physiological model. Toxicol Let 43:117-137. Faustman, EM; Lewandowski, TA; Ponce, RA; et al. (1999) Biologically based dose-response models for developmental toxicants: lessons from methylmercury. Inhal Toxicol 11:559-572. Fennell, TR; Sumner, SC; Waler, VE. (1992) A model for the formation and removal of hemoglobin adducts. Cancer Epidemiol Biomarkers Prev 1:213-219. Filser, JG; Bolt, HM. (1979) Pharmacokinetics of halogenated ethylenes in rats. Arch Toxicol 42:123-136. Filser, JG; Bolt, HM. (1981) Inhalation pharmacokinetics based on gas uptake studies. I. improvement of kinetic models. Arch Toxicol 47:279-292. Fiserova-Bergerova, V. (1995) Extrapolation of physiological parameters for physiologically based simulation models. Toxicol Lett 79:77-86. Fiserova-Bergerova, V; Diaz, ML. (1986) Determination and prediction of tissue-gas partition coefficients. Int Arch Occup Environ Health 58:75-87. Fisher, JW; Allen, BC. (1993) Evaluating the risk of liver cancer in human exposed to trichloroethylene using physiological models. Risk Anal 13:87-95. 6/29/05 R-5 DRAFT—DO NOT CITE OR QUOTE ------- Fisher, J; Mahle, D; Bankston, L; Greene, R; et al. (1997) Lactational transfer of volatile chemicals in breast milk. AIHAJ 58:425-431. Frederick, CB; Potter, DW; Chang-Mateu, MI; et al. (1992) A physiologically-based pharmacokinetic and pharmacodynamic model to describe the oral dosing of rats with ethyl acrylate and its implications for risk assessment Toxicol Appl Pharmacol 114:246-260. Frederick, CB; Lomax, LG; Black, KA; et al. (2002) Use of a hybrid computational fluid dynamics and physiologically based inhalation model for interspecies dosimetry comparisons of ester vapors. Toxicol Appl Pharmacol 183:23-40. Gabrielsson, J; Bondesson, U. (1987) Constant-rate infusion of nicotine and cotinine. I. a physiological pharmacokinetic analysis of the cotinine disposition, and effects on clearance and distribution in the rat. J Pharmacokinet Biopharm 15:583-599. Gallo, JM; Lam, FC; Perrier, DG. (1987) Area method for the estimation of partition coefficients for physiological pharmacokinetic models. J Pharmacokinet Biopharm 15:271-280. Gargas, ML; Andersen, ME; Clewell, HJ. (1986) A physiologically-based simulation approach for determining metabolic rate constants from gas uptake data. Toxicol Appl Pharmacol 86:341-352. Gargas, ML; Burgess, RJ; Voisard, DE; etal. (1989) Partition coefficients of low molecular weight volatile chemicals in various liquids and tissues. Toxicol Appl Pharmacol 98:87-99. Gearhart, JM; Jepson, GW; Clewell, HJ, III; et al. (1990) Physiologically based pharmacokinetic and pharmacodynamic model for the inhibition of acetylcholinesterase by diisopropylfluorophosphate. Toxicol Appl Pharmacol 106:295-310. Gearhart, JM; Jepson, GW; Clewell, HJ; et al. (1994) Physiologically based pharmacokinetic model for the inhibition of acetylcholinesterase by organophosphate esters. Environ Health Perspect 102:51-60. Gearhart, JM; Clewell, HJ; Crump, K.S; et al. (1995) Pharmacokinetic dose estimates of mercury in children and dose-response curves of performance tests in a large epidemiological study. Water Air and Soil Pollut 80:49-58. Gelman, A; Bois, F; Jiang, J. (1996) Physiological pharmacokinetic analysis using population modeling and informative prior distributions. J Amer Stat Assoc 91:436. Gentry, PR; Covington, TR; Andersen, ME; et al. (2002) Application of a physiologically based pharmacokinetic model for isopropanol in the derivation of a reference dose and reference concentration. Reg Toxicol Pharmacol 36:51-68. Gentry, PR; Covington, TR; Clewell, HJ, III. (2003) Evaluation of the potential impact of pharmacokinetic differences on tissue dosimetry in offspring during pregnancy and lactation. Reg Toxicol Pharmacol 38:1-16. Georgopoulos, PG; Roy, A; Gallo, MA. (1994) Reconstruction of short-term multiroute exposure to volatile organic compounds using physiologically based pharmacokinetic models. J Exp Anal Environ Epidemiol 4:309-328. Gerrity, TR; Henry, CJ; Bimbaum, L. (1990) Principles of route-to-route extrapolation for risk assessment. New York, NY: Elsevier. Ginsberg, G; Hattis, D; Russ, A; et al. (2004) Physiologically based pharmacokinetic (PBPK) modeling of caffeine and theophylline in neonates and adults: implications for assessing children's risks from environmental agents. J Toxicol Environ Health A 67:297-329. Haber, LT; Maier, A; Gentry, PR; et al. (2002) Genetic polymorphisms in assessing interindividual variability in delivered dose. Reg Toxicol Pharmacol 35:177-197. 6/29/05 R-6 DRAFT—DO NOT CITE OR QUOTE ------- Haddad, S; Gad, SC; Tardif, R; et al. (1995) Statistical approaches for the validation of physiologically-based pharmacokinetic (PBPK) models. The lexicologist 15:48. (Abstract) Haddad, S; Pelekis, M; Krishnan. K. (1996) A methodology for solving physiologically based pharmacokinetic models without the use of simulation softwares. Toxicol Lett 85:113-126. Haddad, S; Withey, JR; Tardif, R; et al. (1997) Determination of the rate of pyrene metabolism in rat liver post- mitochondrial fractions. Toxicol Lett 93:177-184. Haddad, S; Withey, JR; Lapare, S; et al. (1998) Physiologically based pharmacokinetic modeling of pyrene in the rat. Environ Toxicol Pharmacol 5:245-255. Haddad, S; Restieri, C; Krishnan, K. (2001 a) Characterization of age-related changes in body weight and organ weights from birth to adolescence in humans. J Toxicol Environ Health A 64:453^464. Haddad, S; Beliveau, M; Tardif, R; et al. (2001b) A PBPK modeling-based approach to account for interactions in the health risk assessment of chemical mixtures. Toxicol Sci 63:125-131. Hanna, LM; Lou, SR. (2001) Mass transport analysis: inhalation RFC methods framework for interspecies dosimetric adjustment. Inhal Toxicol 13:437—463. Hattis, D; White, P; Marmorstein, L; et al. (1990) Uncertainties in pharmacokinetics modeling for perchloroethylene. I. comparison of model structure, parameters, and predictions for low dose metabolic rates for models by different authors. Risk Anal 10:449-458. Hattis, D; Banati. P; Goble. R: et al. (1999) Human interindividual variability in parameter's related to health risks. Risk Anal 19:711-726. Hattis, D; Ginsberg, G; Sonawane, B; et al. (2003) Differences in pharmacokinetics between children and adults. II. children's variability in drug elimination half-lives and in some parameters needed for physiologically-based pharmacokinetic modeling. Risk Anal 23:117-142. Hetrick, DM; Jarabek, AM; Travis, CC. (1991) Sensitivity analysis for physiologically-based pharmacokinetic models. J Pharmacokinet Biopharm 19:1 -20. Himmelstein, KJ; Lute, RJ. (1979) A review of the application of physiologically based pharmacokinetic modeling. J Pharm Biopharm 7:127-145. Hissink, EM; Bogaards, JJP; Freidig, AP; et al. (2002) The use of in vitro metabolic parameters and physiologically based pharmacokinetic (PBPK) modeling to explore the risk assessment of trichloroethylene. Environ Toxicol Pharmacol 11:259-271. Hoang, KCT. (1995) Physiologically based pharmacokinetic models-mathematical fundamentals and simulation implementations. Toxicol Lett 79:87-98. Holmes, SL; Ward, RC; Galambos, JD; et al. (2000) A method for optimization of pharmacokinetic models. Toxicol Methods 10:41-53. Horton, VL; Higuchi, MA; Rickert, DE. (1992) Physiologically based pharmacokinetic model for methanol in rats, monkeys and humans. Toxicol Appl Pharmacol 117:26-36. Hurst, CH; DeVito, MJ; Setzer, RW; et al. (2000) Acute administration of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)in pregnant Long Evans rats: association of measured tissue concentrations with developmental effects. Toxicol Sci 53:411-420. 6/29/05 R-7 DRAFT—DO NOT CITE OR QUOTE ------- Hwang, IY; Reardon, KF; Tessari, JD; et al. (1996) A gas-liquid system for enzyme kinetic studies of volatile organic chemicals: determination of enzyme kinetic constants and partition coefficients of trichloroethylene. Drug Metabol Dispos 24:377-382. Igari, Y; Sugiyama, Y; Sawada, Y; et al. (1983) Prediction of diazepam disposition in rat and man by a physiologically-based pharmacokinetic model. J Pharmaco Biopharm 11:577-593. IPCS (International Programme on Chemical Safety. (2001) Guidance document for the use of data in development of chemical-specific adjustment factors (CSAFs) for interspecies differences and human variability in dose/concentration-response assessment. International Programme on Chemical Safety. Geneva, Switzerland: World Health Organization. Isukapalli, SS; Roy, A; Georgopoulos, PG. (1998) Stochastic response surface methods (SRSMs) for uncertainty propagation: application to environmental and biological systems. Risk Anal 18:351-363. Iwatsubo, T; Hiriko, N; Ooie, T; et al. (1996) Prediction of in vivo drug disposition from in vitro data based on physiological pharmacokinetics. Biopharmaceut Drug Dispos 17:273-310. Jarabek, AM. (1994) Inhalation RfC methodology: dosimetric adjustments and dose-response estimation of non- cancer toxicity in the upper respiratory tract. Inhal Toxicol 6:301-325. Jarabek, AM. (1995a) Interspecies extrapolation based on mechanistic determinants of chemical disposition. Human Ecol Risk Assess 1:641-662. Jarabek, AM. (1995b) The application of dosimetry models to identify key processes and parameters for default dose-response assessment approaches. Toxicol Lett 79:171-184. Jarabek, AM; Fisher, J W; Rubenstein, R; et al. (1994) Mechanistic insights aid the search for CFC substitutes: risk assessment of HCFC-123 as an example. Risk Anal 14:231-250. Jepson, GW; Hoover, DK; Black, RK; et al. (1994) A partition coefficient determination method for nonvolatile chemicals in biological tissues. Fund Applied Toxicol 22:519-524. Johanson, G; Dynesius, B. (1988) Liquid: air partition coefficients for six commonly used glycol ethers. Brit J IndustMed 45:561-564. Jonsson, F; Johanson, G. (2001) Bayesian estimation of variability in adipose tissue blood flow in man by physiologically based pharmacokinetic modeling of inhalation exposure to toluene. Toxicol 157:177-193. Jonsson, F; Johanson, G. (2002) Physiologically based modeling of the inhalation kinetics of styrene in humans using a Bayesian population approach. Toxicol Appl Pharmacol 179:35-49. Jonsson, F; Johanson, G. (2003) The Bayesian population approach to physiological toxicokinetic-toxicodynamic models—an example using the MCSim software. Toxicol Lett 143-150. Johanson G. (1991) Modelling of respiratory exchange of polar solvents. Ann Occup Hyg 35:323-39. Johanson, G; Naslund, PH. (1988) Spreadsheet programming: a new approach in physiologically based modeling of solvent toxicokinetics. Toxicol Lett 41:115-127. Johanson, G; Jonsson, F; Bois, F. (1999) Development of new technique for risk assessment using physiologically based toxicokinetic models. Am J Ind Med 36(Suppl 1):101-103. Kaneko, T; Wang, PY; Sato, A. (1994) Partition coefficients of some acetate esters and alcohols in water, blood, olive oil, and rat tissues. Occup Environ Med 51:68-72. 6/29/05 R-8 DRAFT—DO NOT CITE OR QUOTE ------- Karba, R; Zupancic, B; Bremsak, F. (1990) Simulation tools in pharmacokinetic modelling. Acta Pharm Jugosl 40:247-262. Kedderis, GL; Lipscomb, JC. (2001) Application of in vitro biotransformation data and pharmacokinetic modeling to risk assessment. Toxicol and Indust Health 17:315-3 21. Kedderis, GL; Held, SD. (1996) Prediction of furan pharmacokinetics from hepatocyte studies: comparison of bioactivation and hepatic dosimetry in rats, mice, and humans. Toxicol Appl Pharmacol 140:124-130. Kenyon, EM; Kraichely, RE; Hudson, KT; et al. (1996) Differences in rates of benzene metabolism correlate with observed genotoxicity. Toxicol Appl Pharmacol 136:49-56. Keys, DA; Bruckner, JV; Muralidhara, S; et al. (2003) Tissue dosimetry expansion and cross-validation of rat and mice physiologically-based pharmacokinetic models for trichloroethylene. Toxicol Sci 76:35-50. Kim, AH; Kohn, MC; Portier, CJ; et al. (2002) Impact of physiologically based pharmacokinetic modeling on benchmark dose calculations for TCDD-induced biochemical responses. Regul Toxicol Pharmacol 36:287-296. Kimbell, JS; Gross, EA; Joyner, DR; et al. (1993) Application of computational fluid dynamics to regional dosimetry of inhaled chemicals in the upper respiratory tract of the rat. Toxicol Appl Pharmacol 121:253-263. Kirman, CR; Hays, SM; Kedderis, GL; et al. (2000) Improving cancer dose-response characterization by using physiologically based pharmacokinetic modeling: an analysis of pooled data for acrylonitrile-induced brain tumors to assess cancer potency in the rat. Risk Anal 20:13 5—151. Kohn, MC. (1995) Achieving credibility in risk assessment models. Toxicol Lett 79:107-114. Kohn, MC; Melnick, RL. (1999) A physiological model for ligand-induced accumulation of alpha 2u globulin in male rat kidney: roles of protein synthesis and lysosomal degradation in the renal dosimetry of 2,4,4-trimethyl-2- pentanol. Toxicology 136:89-105. Kratochwil, NA; Huber, W; Muller, F; et al. (2002) Predicting plasma protein binding of drugs: a new approach. Biochem Pharmacol 64:1355-1374. Krewski, D; Wang, Y; Bartlett, S; et al. (1995) Uncertainty, variability, and sensitivity analysis in physiological pharmacokinetic models. J Biopharm Stat 5:245-271. Krishnan, K; Andersen, M. (1998) Physiologically based pharmacokinetic models in the risk assessment of developmental neurotoxicants. Slikker, W; ed. In: Handbook of developmental neurotoxicology. New York: Academic Press; pp. 709-725. Krishnan, K; Andersen, ME. (2001) Physiologically based pharmacokinetic modeling in toxicology. Hayes, AW; ed. In: Principles and methods of toxicology. Philadelphia, PA: Taylor & Francis; pp. 193-241. Krishnan, K; Gargas, ML; Fennell, TR; et al. (1992) A physiologically based description of ethylene oxide dosimetry in the rat. Toxicol and Indust Health 8:121-140. Kumagai, S; Matsunaga, I. (1995) Physiologically based pharmacokinetic model for acetone. Occup Environ Med 52:344-352. Lam, G; Chen, ML; Chiou, WL. (1982) Determination of tissue:blood partition coefficients in physiologically-based pharmacokinetic models. J Pharmaceut Sci 71:454-456. Leggett, RW; Williams, LR. (1991) Suggested reference values for regional blood volumes in humans. Health Physics 60:139-154. 6/29/05 R-9 DRAFT—DO NOT CITE OR QUOTE ------- Leroux, BG; Leisenring, WM; Moolgavkar, SH; et al. (1996) A biologically-based dose-response model for development toxicology. Risk Anal 16(4):449—458. Leung, HW. (1991) Development and utilization of physiologically based pharmacokinetic models for toxicological applications. J Toxicol Environ Health 32:247-267. Leung, HW. (1992) Use of physiologically based pharmacokinetic models to establish biological exposure indexes. Am Ind Hyg Assoc J 53:369-374. Leung, HW; Paustenbach, DJ. (1990) Cancer risk assessment for dioxane based upon a physiologically-based pharmacokinetic modeling approach. Toxicol Lett 51:147-162. Levesque, B; Ayotte, P; Tardif, R; et al. (2000) Evaluation of the health risk associated with exposure to chloroform in indoor swimming pools. J Toxicol Environ Health 61:225-243. Levesque, B; Ayotte, P; Tardif, R; et al. (2002) Cancer risk associated with household exposure to chloroform. J Toxicol Environ Health 56:489-502. Lill, MA; Vedani, A; Dobler, M. (2004) Raptor: combining dual-shell representation, induced-fit simulation, and hydrophobicity scoring in receptor modeling: application toward the simulation of structurally diverse ligand sets. J. Med. Chem. 47, 6174-6186. Lin, JH; Sugiyama, Y; Awazu, S; et al. (1982) In vitro and in vivo evaluation of the tissue to blood partition coefficients for physiological pharmacokinetic models. J Pharmacokin Biopharm 10:637-647. Lindstedt, SL; Schaeffer; PJ. (2002) Use of allometry in predicting anatomical and physiological parameters of mammals. Lab. Anim. 36,1-19. Lipscomb, JC; Kedderis, GL. (2002) Incorporating human interindividual biotransformation variance in health risk assessment. Sci Tot Environ 288:13-21. Lipscomb, JC; Fisher, JW; Confer, PD; et al. (1998) In vitro to in vivo extrapolation for trichloroethylene metabolism in humans. Toxicol Appl Pharmacol 152:376-387. Lipscomb, JC; Teuschler, LK; Swartout, J; et al. (2003) The impact of cytochrome P450 2E1-dependent metabolic variance on a risk-relevant pharmacokinetic outcome in humans. Risk Anal 6:1221-1238. Luebeck, EG; Moolgavkar, SH; Buchmann, A; et al. (1991) Effects of polychlorinated biphenyls in rat liver: quantitative analysis of enzyme-altered foci. Toxicol Appl Pharmacol 111 :469—484. Luecke, RH; Wosilait, WD; Pearce, BA; et al. (1994) A physiological based pharmacokinetic computer model for human pregnancy. Teratology 49:90-103. Luecke, RH; Wosilait, WD; Young, JF. (1997) Mathematical analysis for teratogenic sensitivity. Teratology 55:373-380. MacDonald, AJ; Rostami-Hodjegan, A; Tucker, GT; et al. (2002) Analysis of solvent central nervous system toxicity and ethanol interactions using a human population physiologically based kinetic and dynamic model. Regul Toxicol Pharmacol 35(2Pt 1):165-176. Martonen, TB; Zhang, Z; Yu, G; et al. (2001) Three-dimensional computer modeling of the human upper respiratory tract. Cell Biochem Biophys 35:255—61. Medinsky, MA; Kimbell, JS; Morris, JB; et al. (1993) Advances in biologically based models for respiratory tract uptake of inhaled volatiles. Fundam Appl Toxicol 20:265-272. 6/29/05 R-10 DRAFT—DO NOT CITE OR QUOTE ------- Meek, ME; Beauchamp, R; Long, G; et al. (2002) Chloroform: exposure estimation, hazard characterization, and exposure-response analysis. J Toxicol Environ Health B Grit Rev 5:283-334. Melnick, RL; Kohn, MC. (2000) Dose-response analyses of experimental cancer data. Drug Metab Rev 32:193- 209. Menzel, DB; Wolpert, RL; Boger, JR; et al. (1987) Resources available for simulation in toxicology: specialized computers, generalized software and communication networks. Drink Water and Health 8:229-254. Monro, A. (1994) Drug toxicokinetics: scope and limitations that arise from species differences in pharmacodynamic and carcinogenic responses. J Pharmaceut Biopharm 22:41-57. Moolgavkar, S; Knudson, A. (1981) Mutation and cancer: a model for human carcinogenesis. J Natl Cancer Inst 66:1037-1052. Moolgavkar, SH; Luebeck, G. (1990) Two-event model for carcinogenesis: biological, mathematical, and statistical considerations. Risk Anal 10:323-341. Moolgavkar, S; Venzon, D. (1979) Two-event models for carcinogenesis: Incidence curve for childhood and adult tumors. Math Biosci 47:55-77. Mortensen, B; Nilsen, OG. (1998) Allometric species comparison of toluene and n-hexane metabolism: Prediction of hepatic clearance in man from experiments with rodent liver S9 inheadspace vial equilibration system. Pharmacol Toxicol 82:183-188. Mortensen, B; Lokken, T; Zahlsen, K; et al. (1997) Comparison and in vivo relevance of two different in vitro headspace metabolic systems: liver S9 and liver slices. Pharmacol Toxicol 81:35-41. Murphy, JE; Janszen, DB; Gargas, ML. (1995) An in vitro method for determination of tissue partition coefficients of non-volatile chemicals such as 2,3,7,8-tetrachlorodibenzo-p-dioxin and estradiol. J Appl Toxicol 15:147-152. Naumann, BD; Silverman, KC; Dixit, R; et al. (2001) Case studies of categorical data-derived adjustment factors. Human Ecol Risk Assess 7:61-105. Nestorov, IA. (1998) A WWW resource for physiologically based modelling in pharmacokinetics, pharmacodynamics, toxicology and risk assessment. Med Inform (Lond) 23:193-198. Nestorov, I A. (2001) Modelling and simulation of variability and uncertainty in toxicokinetics and pharmacokinetics. Toxicol Lett 120:411—420. Nichols, J; Rheingans, P; Lothenbach, D; et al. (1994) Three-dimensional visualization of physiologically based kinetic model outputs. Environ Health Perspect 102:952-956. O'Flaherty, EJ (1981) Toxicants and Drugs: Kinetics and Dynamics. John Wiley & Sons, New York. O'Flaherty, EJ. (1994) Physiologic changes during growth and development. Environmental Health Perspect 102:103-106. Overton, JH. (2001) Dosimetry modeling of highly soluble reactive gases in the respiratory tract. Inhal Toxicol 13:347-357. Overton, JH; Kimbell, JS; Miller, FJ. (2001) Dosimetry modeling of inhaled formaldehyde: the human respiratory tract. Toxicol Sci 64:122-134. Page, NP; Singh, DV; Farland, W; et al. (1997) Implementation of EPA revised cancer assessment guidelines: incorporation of mechanistic and pharmacokinetic data. Fundamen Appl Toxicol 37:16-36. 6/29/05 R-l 1 DRAFT—DO NOT CITE OR QUOTE ------- Pauluhn, J. (2003) Issues of dosimetry in inhalation toxicity. Toxicol Lett 140:229-238. Paustenbach, DJ. (2000) The practice of exposure assessment: a state-of-art review. J Tox Environ Health (Part B) 3:179-291. Payne, MP; Kenny, LC. (2002) Comparison of models for the estimation of biological partition coefficients. J Toxicol Environ Health 65:897-931. Pelekis, M; Nicolich, MJ; Gauthier, JS. (2003) Probabilistic framework for the estimation of the adult and child toxicokinetic intraspecies uncertainty factors. Risk Anal 23:1239-1255. Perbellini, L; Mozzo, P; Olivata, D; et al. (1990) Dynamic biological exposure indexes for n-hexane and 2,5- hexanedione, suggested by a physiologically-based pharmacokinetic model. American Indust Hygiene Assoc J 51:356-362. Perkins RA, Ward KW, Pollack GM. (1995) A pharmacokinetic model of inhaled methanol in humans and comparison to methanol disposition in mice and rats. Environ Health Perspect. 103:726-33 Pierce, CH; Dills, RL; Morgan, MS; et al. (1998) Biological monitoring of controlled toluene exposure. Int Arch Occup Environ Health 71:43 3-444. Ploemen, JP; Wormhoudt, LW; Haenen, GR; et al. (1997) The use of human in vitro metabolic parameters to explore the risk assessment of hazardous compounds: the case of ethylene dibromide. Toxicol Appl Pharmacol 143:56-69. Poet, TS; Soelberg, JJ; Weitz, KK; et al. (2003) Mode of action and pharmacokinetic studies of 2-butoxyethanol in the mouse with an emphasis on forestomach dosimetry. Toxicol Sci 71:176-189. Portier, CJ; Kaplan, NL. (1989) Variability of safe dose estimates when using complicated models of the carcinogenic process. A case study: methylene chloride. Fundam Appl Toxicol 13:533-44. Portier, C; Tritscher, A; Kohn, M; et al. (1993) Ligand/receptor binding for 2,3,7,8-TCDD: implications for risk assessment. Fundam Appl Toxicol 20:48-56. Poulin, P; Krishnan, K. (1995) A biologically-based algorithm for predicting human tissue: blood partition coefficients of organic chemicals. Hum Exp Toxicol 14:273-280. Poulin, P; Krishnan, K. (1996a) A tissue composition-based algorithm for predicting tissue: air partition coefficients of organic chemicals. Toxicol Appl Pharmacol 136:126-130. Poulin, P; Krishnan, K. (1996b) A mechanistic algorithm for predicting blood:air partition coefficients of organic chemicals with the consideration of reversible binding in hemoglobin. Toxicol Appl Pharmacol 136:131-137. Poulin, P; Theil, FP. (2000) A priori prediction of tissue:plasma partition coefficients of drugs to facilitate the use of physiologically-based pharmacokinetic models in drug discovery. J Pharma Sci 89:16-35. Price, PS; Conelly, RB; Chaisson, CF; et al. (2003a) Modeling interindividual variation in physiological factors used in PBPK models of humans. Grit Rev in Toxicol 33:469-503. Price, K; Haddad, S; Krishnan, K. (2003b) Physiological modeling of age-specific changes in the pharmacokinetics of organic chemicals in children. J Toxicol Environ Health 66:417-433. Ramsey, JC; Andersen, ME. (1984) A physiologically-based description of the inhalation pharmacokinetics of styrene in rats and humans. Toxicol Appl Pharmacol 73:159-175. Rao, HV; Ginsberg, GL. (1997) A physiologically-based pharmacokinetic model assessment of methyl t-butyl ether in groundwater for a bathing and showering determination. Risk Anal 17:583-598. 6/29/05 R-l 2 DRAFT—DO NOT CITE OR QUOTE ------- Reddy, MB; Andersen, ME; Morrow, PE; et al. (2003) Physiological modeling of inhalation kinetics of octamethylcyclotetrasiloxane in humans during rest and exercise. Toxicol Sci 72:3-18. Reitz, RH; McDougal, JN; Himmelstein, MW; et al. (1988a) Physiologically-based pharmacokinetic modeling with methyl chloroform: implications for interspecies, high-low dose and dose-route extrapolations. Toxicol Appl Pharmacol 95:185-199. Reitz, RH; Mandrela, AL; Park, CN; et al. (1988b) Incorporation of in vitro enzyme data into the physiologically- based pharmacokinetic (PBPK) model for methylene chloride: implications for risk assessment. Toxicol Let 43:97- 116. Reitz, RH; Mandrela, AL; Guengerich, FP. (1989) In vitro metabolism of methylene chloride in human and animal tissues: use in physiologically-based pharmacokinetic models. Toxicol Appl Pharmacol 97:230-246. Reitz, RH; Mandrela, AL; Corley, RA; et al. (1990a) Estimating the risk of liver cancer associated with human exposures to chloroform using physiologically-based pharmacokinetic modeling. Toxicol Appl Pharmacol 105:443^59. Reitz, RH; McCroskey, PS; Park, CN; et al. (1990b) Development of a physiologically-based pharmacokinetic model for risk assessment with 1,4-dioxane. Toxicol Appl Pharmacol 105:37-54. Reitz, RH; Gargas, ML; Mendrala, AL; etal. (1996a) In vivo and in vitro studies of perchloroethylene metabolism for physiologically based pharmacokinetic modeling in rats, mice, and humans. Toxicol Appl Pharmacol 136:289- 306. Reitz, RH; Gargas, ML; Andersen, ME; et ai. (1996b) Predicting cancer risk from vinyl chloride exposure with a physiologically based pharmacokinetic model. Toxicol Appl Pharmacol 137:253-267. Renwick, AG. (2001) Toxicokinetics-pharmacokinetics in toxicology. Hayes, WA, ed. In: Principles and methods of toxicology, 4th Edition. Philadelphia: Taylor & Francis; pp. 137-192. Renwick, AG; Lazarus, NR. (1998) Human variability and noncancer risk assessment: an analysis of the default uncertainty factor. Regul Toxicol Pharmacol 27:3-20. Rey, TD; Havranek, WA. (1996) Some aspects of using the SimuSolv program for environmental, pharmacokinetics and toxicological applications. Ecological Modeling 86:277-282. Rogers, JM; Mole, ML; Chernodd, N; et al. (1993) The developmental toxicity of inhaled methanol in the CD-I mouse, with quantitative dose-response modeling for estimation of benchmark doses. Teratology 47:175-188. Rowland, M. (1985) Physiologic pharmacokinetic models and interanimal species scaling. Pharmacol Ther 29:49- 68. Roy, A; Georgopoulos, PG. (1998) Reconstructing week-long exposures to volatile organic compounds using physiologically based pharmacokinetic models. J Expo Anal Environ Epidemiol 8:407-422. Roy, A; Weisel, CP; Lioy, PJ; et al. (1996) A distributed parameter physiologically-based pharmacokinetic model for dermal and inhalation exposure to volatile organic compounds. Risk Anal 16:147-160. Santostefano, MJ; Wang, X; Richardson, VM; et al. (1998) A pharmacodynamic analysis of TCDD-induced cytochrome P450 gene expression in multiple tissues: dose- and time-dependent effects. Toxicol Appl Pharmacol 151:294-310. Sato, A; Nakajima, T. (1979) Partition coefficients of some aromatic hydrocarbons and ketones in water, blood and oil. Brit J Indust Med 36:231-234. 6/29/05 R-l3 DRAFT—DO NOT CITE OR QUOTE ------- Sato, A; Endoh, K; Kaneko, T; et al. (1991) A simulation study of physiological factors affecting phamacokinetic behavior of organic solvent vapors. Brit J Indust Med 48:342-347. Schlosser, PM; Patrick, DL; Conolly, RB; et al. (2003) Benchmark dose risk assessment for formaldehyde using airflow modeling and a single-compartment, DNA-protein cross-link dosimetry model to estimate human equivalent doses. Risk Anal 23:473-487. Schoeffner, DJ; Warren, DA; Muralidhara, S; et al. (1999) Organ weights and fat volume in rats as a function of strain and age. J Toxicol Environ Health 56:449—462. Silverman, KG; Naumann, BD; Holder, DJ; et al. (1999) Establishing data-derived adjustment factors from published pharmaceutical clinical trial data. Hum Ecol Risk Assess 5:1059-1089. Smith, AE; Gray, GM; Evans, JS. (1995) The ability of predicted internal dose measures to reconcile tumor bioassay data for chloroform. Regul Toxicol Pharmacol 21:339-351. Steinbach, KH; Raffler, H; Pabst, G; et al. (1980) A mathematical model of canine granulocytopoiesis. J Math Biol 10:1-12. Sultatos, LG; Kim, B; Woods, L. (1990) Evaluation of estimations in vitro of tissue: blood distribution coefficients for organothiophosphate insecticides. Toxicol Appl Pharmacol 103:52-55. Sweeney, LM; Tyler, TR; Kirman, CR; et al. (2001) Proposed occupational exposure limits for select ethylene glycol ethers using PBPK models and Monte Carlo simulations. Toxicol Sci62(l):124-139. Tan, YM; Butterworth, BE; Gargas, ML; et al. (2003) Biologically motivated computational modeling of chloroform cytolethality and regenerative cellular proliferation. Toxicol Sci 75 :192-200. Tardif, R; Charest-Tardif, G; Brodeur, J; et al. (1997) Physiologically based pharmacokinetic modeling of a ternary mixture of alkyl benzenes in rats and humans. Toxicol Appl Pharmacol 144:120-134. Terasaki, T;Iga, T; Sugiyama, Y; etal. (1984) Nuclear binding as a determinant of tissue distribution of adriomycin, daunomycin, adriamycinol, daunorubicinol and actinomycin D. J Pharmacobio-Dynamics 7:269-277. Thomas, RS; Yang, RSH; Morgan, DG; et al. (1996a). PBPK modeling/Monte Carlo simulation of methylene chloride kinetic changes in mice in relation to age and acute, subchronic, and chronic inhalation exposure. Environ Health Perspect 104:858-865. Thomas, RS; Lytle, WE; Keefe, TJ; et al. (1996b) Incorporating Monte Carlo simulation into physiologically based pharmacokinetic models using advanced continuous simulation language (ACSL): a computational method. Fundam Appl Pharmacol 31:19-28. Thomas, RS; Conolly, RB; Gustafson, DL; et al. (2000) A physiologically based pharmacodynamic analysis of hepatic foci within a medium-term liver bioassay using pentachlorobenzene as a promoter and diethylnitrosamine as an initiator. Toxicol Appl Pharmacol 166:128-137. Thrall, KD; Vucelick, ME; Gies, RA; et al. (2000). Comparative metabolism of carbon tetrachloride in rats, mice, and hamsters using gas uptake and PBPK modeling. J Toxicol Environ Health 60:531-548. Timchalk, C; Poet, TS; Lin, Y; et al. (2001) Development of an integrated microanalytical system for analysis of lead in saliva and linkage to a physiologically based pharmacokinetic model describing lead saliva secretion. AIHAJ 62:295-302. Timchalk, C; Nolan, RJ; Mendrala, AL; et al. (2002) A physiologically based pharmacokinetic and pharmacodynamic (PBPK/PD) model for the organophosphate insecticide chlorpyrifos in rats and humans. Toxicol Sci 66:34-53. 6/29/05 R-14 DRAFT—DO NOT CITE OR QUOTE ------- Tran, CL; Jones, AD; Cullen, RT; et al. (1999) Mathematical modeling of the retention and clearance of low-toxicity particles in the lung. InhToxicol 11:1059-1076. Travis, CC; Hattemer-Frey, HA. (1991) Physiological pharmacokinetic models. Krewski, D; Franklin, C; eds. In: Statistics in Toxicology. New York: Gordon and Breach; p. 170. U.S. EPA (Environmental Protection Agency). (1991) Risk assessment guidance for superfund sites: volume III. Part A: Process for conducting probabilistic risk assessment. Office of Emergency and Remedial Response, Washington, DC; EPA 540-R-02-002. U. S. EPA (Environmental Protection Agency). (1992) Guidelines for exposure assessment. Federal Register 57(104):22888-22938. Available online at http://www.epa.gov/ncea/raf/rafguid.htm. U. S. EPA. (Environmental Protection Agency). (1994). Methods for derivation of inhalation reference concentrations and application of inhalation dosimetry. Office of Research and Development, Washington, DC; EPA/600/8-90/066F. U.S. EPA (Environmental Protection Agency). (1997) Guiding principles for Monte Carlo analysis. Risk Assessment Forum, Washington, DC; EPA/630/R-97/001. U.S. EPA (Environmental Protection Agency). (1999a) Extrapolation of the benzene inhalation unit risk estimate to the oral route of exposure [draft]. National Center for Environmental Assessment, Washington, DC; NCEA-W- 0517. U.S. EPA (Environmental Protection Agency). (1999b) Toxicological review of ethylene glycol monobutyl ether. In support of summary information on the IRIS. National Center for Environmental Assessment, Washington, DC. Available online at http://www.epa.gov/iris. U.S. EPA (Environmental Protection Agency). (2000a) Benchmark dose technical guidance document [external review draft]. Risk Assessment Forum, Washington, DC; EPA/630/R-00/001. U.S. EPA (Environmental Protection Agency). (2000b) Toxicological review of vinyl chloride. In support of summary information on the IRIS. National Center for Environmental Assessment, Washington, DC; EPA/635/R- 00/004. Available online at http://www.epa.gov/iris. U.S. EPA (Environmental Protection Agency). (2002) A review of the reference dose and reference concentration processes. Risk Assessment Forum, Washington, DC; EPA/630/P-02/002-F. Available online at http ://epa.gov/ncea/raf. U.S. EPA (Environmental Protection Agency). (2003) Toxicological review of xylenes. In support of summary information on the IRIS, Washington, DC; EPA/635/R-03/001. Available online at http://www.epa.gov.iris. U.S. EPA (Environmental Protection Agency). (2004) Air quality criteria for particulate matter. National Center for Environmental Assessment, Research Triangle Park, NC; EPA/600/P-99/002aF. U.S. EPA (Environmental Protection Agency). (2005a) Guidelines for carcinogen risk assessment. Federal Register 70(66)17765—17817. Available online at http://www.epa.gov/cancerguidelines. U.S. EPA (Environmental Protection Agency). (2005b) Supplemental guidance for assessing susceptibility from early-life exposure to carcinogens. Risk Assessment Forum, Washington, DC; EPA/630/R-03/003F. Van Asperen, J; Rijcken, WRP; Lammers, JHCM. (2003) Application of physiologically based toxicokinetic modelling to study the impact of the exposure scenario on the toxicokinetics and the behavioural effects of toluene in rats. ToxicolLett 138:51-68. Vicini, P; Pierce, CH; Dills, RL; et al. (1999). Individual prior information in a physiological model of 2H8-toluene kinetics: an empirical Bayes estimation strategy. Risk Anal 19:1127-1134. 6/29/05 R-l5 DRAFT—DO NOT CITE OR QUOTE ------- Vinegar, A; Jepson, GW. (1996) Cardiac sensitization thresholds of halon replacement chemicals predicted in humans by physiologically based pharmacokinetic modeling. Risk Anal 16:571-579. Vinegar, A; Winsett, DW; Andersen, ME; et al. (1990) Use of a physiologically based pharmacokinetic and computer simulation for retrospective assessment of exposure to volatile toxicants. Inhal Toxicol 2:119-128. Voisin, E; Ruthsatz, M; Collins, J; et al. (1990) Extrapolation of animal toxicity to humans: interspecies comparisons in drug development. Regul Toxicol Pharmacol 12:107-116. Wagner, JG. (1981) History of pharmacokinetics. Pharmacol Ther 12:537-562. Walton, K; Dome, JL; Renwick, AG. (2001) Uncertainty factors for chemical risk assessment: interspecies differences in the in vivo pharmacokinetics and metabolism of human CYP1A2 substrates. Food Chem Toxicol 39:667-680. Welsch, F; Blumenthal, GM; Conolly, RB. (1995) Physiologically based pharmacokinetic models applicable to organogenesis: extrapolation between species and potential use in prenatal toxicity risk assessments. Toxicol Lett (82-83):539-547. Williams, RJ; Vinegar, A; McDougal, JN; et al. (1996) Rat to human extrapolation of HCFC-123 kinetics deduced from halothane kinetics-a corollary approach to physiologically based pharmacokinetic modelling. Fund Appl Toxicol 30:55-66. Wunscher, G; Kersting, H; Heberer, H; et al. (1991) Simulation system SONCHES-based toxicokinetic model and data bank as a tool in biological monitoring and risk assessment. Sci Total Environ 101:101-109. Yates, FE. (1978) Good manners in good modeling : mathematical models and computer simulations of physiological systems. Am J Physiol 234 :R59-R160. 6/29/05 R-l 6 DRAFT—DO NOT CITE OR QUOTE ------- |