Modeling Human Exposure to Indoor Contaminants - ACS Publications

Jul 19, 2016 - Environment and Resource Sciences, Trent University, Peterborough, Ontario K9L 0G2 Canada. §. ExxonMobil Biomedical ... to the develop...
0 downloads 8 Views 431KB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Article

Modeling human exposure to indoor contaminants: external source to body tissues Eva M. Webster, Hua Qian, Donald Mackay, Rebecca D Christensen, Britta Tietjen, and Rosemary Zaleski Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b00895 • Publication Date (Web): 19 Jul 2016 Downloaded from http://pubs.acs.org on July 19, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 30

Environmental Science & Technology

1

Modeling human exposure to indoor contaminants: external source to body tissues

2 3

Eva M. Webstera*, Hua Qianc, Donald Mackayb, Rebecca D. Christensenb, Britta Tietjenb,d,

4

Rosemary Zaleskic

5 6

a

Carrousel Environmental Modelling Research, Peterborough, Ontario K9H 0E2 Canada

7

b

Environment and Resource Sciences, Trent University, Peterborough, Ontario K9L 0G2

8

Canada

9

c

ExxonMobil Biomedical Sciences, Inc., Annandale, New Jersey, USA. 08801

10

d

Freie Universität Berlin, Biodiversity and Ecological Modeling, D-14195 Berlin, Germany

11

* Corresponding author: Email: [email protected], Tel: 705-748-1011

12 13

Keywords: Indoor, PBPK, fugacity, multimedia model

14

ACS Paragon Plus Environment

1

Environmental Science & Technology

15 16

Page 2 of 30

Abstract Information on human indoor exposure is necessary to assess the potential risk to

17

individuals from many chemicals of interest. Dynamic indoor and human physicologically based

18

pharmacokinetic (PBPK) models of the distribution of non-ionizing, organic chemical

19

concentrations in indoor environments resulting in delivered tissue doses are developed,

20

described and tested. The Indoor model successfully reproduced independently measured,

21

reported time-dependent air concentrations of chloroform released during showering and of 2-

22

butyoxyethanol following use of a volatile surface cleaner. The Indoor model predictions were

23

also comparable to those from a higher tier consumer model (ConsExpo 4.1) for the surface

24

cleaner scenario. The PBPK model successful reproduced observed chloroform exhaled air

25

concentrations resulting from an inhalation exposure. Fugacity based modeling provided a

26

seamless description of the partitioning, fluxes, accumulation and release of the chemical in

27

indoor media and tissues of the exposed subject. This has the potential to assist in health risk

28

assessments, provided that appropriate physical/chemical property, usage characteristics, and

29

toxicological information are available.

30

ACS Paragon Plus Environment

2

Page 3 of 30

31

Environmental Science & Technology

Introduction

32 33

Considerable effort has been devoted to measuring and estimating human exposure to

34

chemical substances in the indoor environment 1-13 and enclosed spaces such as automobiles 14-16

35

and airplane cabins 17-20, because both indoor environmental characteristics and human behavior

36

patterns suggest that in many cases indoor exposures may be more important to human health

37

than outdoor exposures. For example, a given mass of chemical released outdoors will result in

38

lower exposure than the same release in an indoor environment because of reduced dispersion

39

and great proximity in the indoor environment, i.e., enclosed indoor environments have a

40

defined, lower air volume with limited ventilation and higher duration of occupancy.

41 42

One widely used single-compartment model is ConsExpo™ 4.1 21. It was developed to

43

estimate chemical exposure from non-dietary use of consumer products in an indoor environment

44

based on the information on the individual exposed (e.g., body weight, inhalation rate), the

45

product use pattern (e.g., product concentration, use amount and frequency), and the exposure

46

location characteristics (e.g., room volume, ventilation) for all relevant exposure routes

47

(inhalation, dermal and oral). The uptake of the chemical in the model is estimated either by

48

applying a fixed uptake fraction and/or by a diffusive uptake based on the chemical skin

49

permeability for the dermal route.

50 51

Multimedia or multi-compartment models can describe the fate, transport and

52

degradation of chemicals indoors with a view to assessing and reducing exposures. Indoor

53

models describe the movement and fate of organic chemicals in domestic spaces such as homes

ACS Paragon Plus Environment

3

Environmental Science & Technology

54

and offices, address a variety of scenarios, and have differing levels of sophistication, model

55

input requirements and model outputs. Reviews of indoor and exposure models are available

56

elsewhere 22-25, only models key to the development of the present models are highlighted here.

Page 4 of 30

57 58

Fugacity-based models express explicitly the relative equilibrium status of the

59

compartments and simplify the equations. For example, Mackay and Paterson 3 developed a

60

simple, steady state model for assessing exposure in a one-room house with a series of

61

resistances to chemical movement from the source to the outdoor air; Matoba and colleagues

62

developed and applied a series of models 4, 26, 27 to domestic pesticide use in Japan. A significant

63

advance was made by Bennett and Furtaw 1 who developed a fugacity-based dynamic mass

64

balance model, for pesticides used in domestic settings, that included a primary treated zone and

65

a secondary untreated zone, each with air, six size fractions of airborne particulate (dust), carpet,

66

smooth vinyl flooring and wall (painted wallboard) compartments. Building on these works,

67

polyurethane foam (PUF) such as in furniture was added to the model to address chemical

68

entering the room from treated PUF, carpet or electronic devices 13. This model can be used to

69

predict the impact of environmental modifications such as introduction of a new electronic

70

devices or human activities that stir household dust, such as walking, sweeping, vacuuming or

71

from the mechanical action of compression and release of PUF in the routine use of soft

72

furnishings.

73 74

Concentration-based dynamic mass balance models have also been developed and applied to

75

indoor environments. A comprehensive and quantitative treatment of phase equilibria and inter-

76

phase kinetics of a wide variety of semi-volatile indoor organic contaminants has been described

ACS Paragon Plus Environment

4

Page 5 of 30

Environmental Science & Technology

77

by Weschler and Nazaroff 10 based on vapor pressure and octanol-air partition coefficient, KOA to

78

describe air-surface sorption and desorption. This modeling demonstrated that, depending on

79

these properties, contaminant levels may be under kinetic or equilibrium control and

80

characteristic times of air-surface exchange vary greatly. Time-scale estimates were made for

81

two scenarios: a crack and crevice pesticide application and three life stages for product

82

containing plasticizers or flame retardants. Nazaroff 28 combined models and empirical data to

83

show that intake fraction, the pollutant mass inhaled per unit mass released from a source, varies

84

with building-related factors such as ventilation, human factors such as breathing rate and

85

chemical properties such as sorption to surfaces and reactivity, i.e., degradation half-life. These

86

approaches estimate the exposure of humans to substances. However, to assess impacts on

87

human health, also pathways into the body and distribution within a body are of importance.

88 89

Physiologically based pharmacokinetic (PBPK) models have long been used to

90

investigate drug distribution within a body 29, 30. Inhalation is commonly modeled as the

91

exposure route but injection, ingestion and dermal sorption are also considered. Blood flow

92

connects tissues or groups of tissues. The grouping of tissues into aggregate compartments has

93

been shown to cause minimal loss of predictive power 31. For example, for acidic and basic drugs

94

the muscle and fat pair was found to be sufficient 32. Most models include liver, fat, muscle and

95

skin; blood may or may not be considered as a compartment. Chemical loss is typically by

96

metabolism in the liver and chemical removal by exhalation and urination. A review of PBPK

97

models is available elsewhere 33.

98

ACS Paragon Plus Environment

5

Environmental Science & Technology

99

Page 6 of 30

Combining knowledge of chemical fate in indoor environments with chemical uptake and

100

kinetics within the body can provide insight into the likely body burdens received by each of the

101

exposure routes of inhalation, ingestion and dermal contact. For toxicological assessments,

102

linking an Indoor and a PBPK human model has the advantage of being able to demonstrate the

103

sources and magnitudes of the chemical emissions, as discussed by Weschler and Nazaroff

104

who included body burden data for their target chemicals. This is most useful if the intake

105

estimates can be converted into tissue concentrations, since it is tissue concentrations that are

106

associated with potential adverse effects. Modeling chemical fate through indoor sources into a

107

human receptor spans nearly three decades from a shower model used to estimate inhalation

108

exposure to volatile organic compounds (VOCs) from water 34 to multimedia indoor models with

109

human intake fractions 35, 36, exposure factors and human activity data 37, and simple PBPK

110

models 12.

10

,

111 112

Goals and Objectives of the present study

113 114

The goal of the present study was to develop and test a transparent and readily applied

115

indoor/human model pair with the indoor model generating a time profile of the air fugacities

116

and concentrations associated with domestic exposure scenarios loosely coupled to a human

117

PBPK model to provide an improved understanding of chemical distribution within the body and

118

internal dose over time. Loose coupling of the models allows the results of the Indoor model to

119

be directly imported into the PBPK model but also allows the models to be used independently,

120

or alternate models to be substituted for either of the pair. The aim is to determine the

ACS Paragon Plus Environment

6

Page 7 of 30

Environmental Science & Technology

121

relationship between individual tissue fugacities and their relative time dependencies as

122

influenced by the exposures by inhalation, ingestion and dermal sorption.

123 124

Model Description

125 126

Overall model structure

127 128

Two individual component models were developed, an indoor environment model and a

129

physiologically-based pharmacokinetic, PBPK, human model. The fugacity concept, briefly

130

described below, was used for both models to provide a seamless description of the chemical

131

path from source to receptor to demonstrate and quantify the source-receptor relationships

132

between the indoor environment and human tissues and their relative time dependencies as

133

influenced by the nature, rates and duration of emissions, chemical degradation, dilution and

134

uptake by inhalation, ingestion and dermal routes. This is accomplished by a loose-coupling

135

mechanism of the models where the time series of fugacities in room air from the Indoor model

136

is stored and becomes an input to the PBPK model. Dermal uptake and ingestion can depend on

137

the Indoor model results but these must be provided to the PBPK model independently, there is

138

no automatic mechanism provided. The component modeling approach provides flexibility in

139

model selection from simple to complex, and further allows the direct substitution of measured

140

observations.

141 142

Only one substance may be addressed at a time. To address mixtures of multiple

143

substances the models must be applied independently for each substance and the results

ACS Paragon Plus Environment

7

Environmental Science & Technology

144

combined; a procedure that is invalid if there are chemical interactions such as may occur in the

145

liver. In such cases individual chemical estimates from the Indoor model can be entered into a

146

multi-substance PBPK model such as that of Haddad et al.38.

Page 8 of 30

147 148

The quantity of chemical absorbed by the human is not included as a loss from indoor air

149

nor does exhaled air contribute to the concentration in indoor air. For chemical originating in the

150

occupied space, this is a conservative assumption. The human is restricted to a single simulated

151

room for the duration of the simulated time. These are issues for future consideration, as

152

discussed later.

153 154

The Fugacity Concept

155 156

Briefly, fugacity is defined as the escaping tendency or partial pressure of a chemical

157

from the phase in which it currently resides. Fugacity, f (Pa), is related to concentration, C

158

(mol/m3), by a fugacity capacity, Z (mol/m3·Pa), such that C equals Z·f. The Z value

159

characterizes the ability of a phase to absorb and retain a chemical and depends on the chemical,

160

the nature of “solvent” phase and temperature. D-values or fugacity rate coefficients are used to

161

express rates of chemical uptake, loss and reaction. These D-values (mol/Pa·h) are essentially

162

fugacity-based transport and transformation rate constants. The chemical fluxes, N (mol/h), by

163

degrading reaction or by transport between phases or compartments are expressed as D·f. D

164

values for advective transport in air are defined as GZA where G is the flow rate (m3/h) and ZA

165

applies to the air phase. Large D-values correspond to fast processes. More detail is available in

166

the textbook by Mackay 39. Conventional concentration or rate constant models are algebraically

ACS Paragon Plus Environment

8

Page 9 of 30

Environmental Science & Technology

167

equivalent to fugacity models but fugacity is convenient for comparing the relative equilibrium

168

status of the chemical in various media.

169 170

Model compartments and processes

171 172

In the Indoor model, a chemical is distributed between the air, wallboard, and hard and

173

soft surfaces in multiple rooms connected by air circulation and with outdoor air exchange, as

174

shown in Figure 1. Details of the fugacity formulation of the Indoor model are given in the

175

Supporting Information (Section S-1). Surfaces are modeled as quasi-liquid layers, QLLs,

176

(Section S-1c-iii). Air-borne and settled particles are included in each modeled as a fraction of

177

the volume of each compartment; a mass balance of particles is not calculated by this model.

178

Time-dependent concentrations of organic chemicals in indoor compartments are distributed by

179

diffusive and particle-associated movement and lost through degradation and ventilation.

180

Chemical in indoor air becomes available for respiratory uptake by a human.

181 182

Human exposure is treated using the multi-compartment PBPK model (Figure 2, Section

183

S-2a). The PBPK model calculates the time-dependent fate and distribution of the chemical

184

between various tissue concentrations to simulate conditions of periodic or continuous exposure.

185

Chemical may enter the body in inhaled air, by dermal absorption, and in ingested food and

186

water. It distributes through the body with blood flow between tissue groups, accumulating or

187

degrading in the fat tissue, richly and poorly perfused tissue, liver, skin and venous and arterial

188

blood, and is removed by excretion, exhalation, dermal desorption, and metabolism. The model

189

allows for the direct entry of measured blood-air partition coefficients or, because these and

ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 30

190

blood-tissue partition coefficients are available for only a limited number of chemicals, they can

191

be calculated in the model from the more widely available KOW values. Determination of blood-

192

tissue, and hence blood-air, partition coefficients from routinely collected physical chemical

193

properties, such as KOW, has been demonstrated to be a valid approach 40, 41. Metabolism or

194

biotransformation of the chemical is treated as first order degradation in each compartment

195

except in the liver where the Michaelis-Menten method is applied, as described in the Supporting

196

Information (Section S-2a-v). PBPK model results are not scalable because of this non-linear

197

treatment of metabolism in the liver. An overall chemical residence time in the human body can

198

deduced as the ratio between the sum of the amounts in all compartments and the sum of intake

199

rates by all routes. This model represents intermediate complexity between the steady state

200

model of Paterson and Mackay 42 and the highly complex, multi-tissue, multi-process, dynamic

201

model of Cahill et al 43.

202 203

Parameter values are not integral to the models; model design allows all of the defining

204

values to be entered by the user. For real systems, when known, actual system-specific

205

experimental values can and should be used. Some parameter values are routinely measured and

206

reported (e.g., room dimensions, the octanol-water partition coefficient, human subject mass),

207

some may need to be estimated from measurements not specific to the space or subject (e.g.,

208

room air exchange rate constant, human subject tissue lipid content or breathing rate), others may

209

be more challenging (e.g., mass transfer coefficients) but, for order of magnitude estimates,

210

generic or default values such as those suggested in the present work may often suffice.

211 212

ACS Paragon Plus Environment

10

Page 11 of 30

213

Environmental Science & Technology

Model Application Scenarios

214 215

Test of model mass balance

216 217

An important requirement of models of this type is that a complete mass balance can be

218

obtained. This is most readily accomplished by running the model to a steady state condition in

219

which total chemical influx must equal total losses of chemical, i.e., there is no net accumulation.

220

This test lends confidence to the reliability of the dynamic model results. For this test a constant

221

600 µg/h emission of an illustrative chemical is considered as released in a small room with an

222

‘Average Man’ over a 120 hour period (5 days). All system defining values are given in Tables

223

S-3a-c. Details of the value selections defining the Average Man are given in the Supporting

224

Information (Table S-2e). An ‘Average Woman’ or “Child” could be defined by the same

225

method but most experiments use adult male subjects.

226 227

Under conditions of normal human activity is such that steady state conditions are unlikely

228

to be achieved; chemical residence time within the body would need to be less than the

229

approximately 8-hour duration of a sleep period, for example, before steady state could

230

reasonably be expected. To ensure that a mass balance is achieved and that all mass fluxes and

231

accumulations are consistent, the PBPK model was tested with the ‘Average Man’ showering

232

naked, described in Table S-3c of the Supporting Information, to the hypothetical, illustrative

233

chemical, described in Table S-3a of the Supporting Information, with constant exposure to, the

234

steady state concentration in air in the test of the Indoor model (9280 ng/m3), an arbitrary

235

concentration in food (3 × 106 ng/kg), and in the dermal contact medium (water; 106 ng/m3). A

ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 30

236

mass balance check was performed on the chemical fluxes after apparent steady state conditions

237

had been achieved.

238 239

Test by comparison: Showering experiments

240 241

The experimental studies of Jo et al. 44 were simulated to test the Indoor and PBPK

242

models. The air in a shower stall was monitored for chloroform concentrations resulting from

243

water use and exhaled air concentrations of chloroform were reported for two scenarios: the

244

participant wearing a rubber suit to prevent dermal sorption, and a normal shower where both

245

dermal sorption and inhalation would occur 44. The properties of chloroform used in the

246

simulations are given in Table S-3a of the Supporting Information. A chemical transfer

247

efficiency (TE = 61%) from the shower water to the air is calculated by the method provided in

248

the Supporting Information (Section S-1c-vi). The Indoor model was parameterized to simulate

249

the experimental laboratory shower stall (Table S-3b in the Supporting Information) and

250

showering scenario (Table S-3d in the Supporting Information). The air exchange was

251

minimized in the experiment by the use of a shower curtain and the absence of an exhaust fan but

252

no value was reported 44. In later experiments using a similar shower chamber under similar test

253

conditions, air exchange was estimated to be between 6.0 and 19.2 air changes per hour 45; a

254

value of 15/h was selected as the model input. Indoor model results were compared to reported

255

air concentrations. The PBPK model was parameterized for the Average Man (Table S-3c in the

256

Supporting Information) because no information was provided on the experiment participants.

257

The area of exposed skin was set to a negligibly small value (10-11 m2) for the ‘inhalation only’

258

experimental scenario and to 1 m2 for the ‘normal showering’ scenario where both inhalation and

ACS Paragon Plus Environment

12

Page 13 of 30

Environmental Science & Technology

259

dermal sorption are considered (Table S-3c in the Supporting Information). The Average Man

260

was exposed to the air concentrations predicted by the Indoor model for reported water

261

concentrations for the ‘inhalation only’ and to both air and water concentrations for the ‘normal

262

showering’ scenario (Table S-3d in the Supporting Information). For convenience in modeling

263

‘normal showering’, during the five minutes of drying after showering the water definition of the

264

dermal contact medium was retained but with a negligible concentration. To be more correct the

265

contact medium should have been redefined as air and the air concentrations applied. PBPK

266

model results were used to generate exhaled air concentrations for comparison to reported

267

observations. The effect of redefining the dermal contact medium for the final five minutes was

268

tested as described in the Supporting Information.

269 270

For simulation of the Xu and Weisel experiments, the Indoor model was parameterized

271

for the experimental shower stall and showering scenario 45, 46 as defined in Tables S-3b and S-

272

3d with chloroform as defined in Table S-3a all of the Supporting Information. Subject C was

273

selected as most similar to the ‘Average Man’ defined in Table S-3c of the Supporting

274

Information. In addition, the reported total blood volume of subject C of 5.5 L (sub-divided

275

0.83:0.17 between venous and arterial, as for ‘Average Man’) 45 was used. The lighter weight of

276

subject C was arbitrarily assigned as a lower fat volume of 0.0150 m3. The area of exposed skin

277

was set to a negligibly small value (10-11) is simulate the experimental condition of inhalation

278

only exposure. All other model inputs for the human subject were as defined for the Average

279

Man in Table S-3c of the Supporting Information.

280 281

ACS Paragon Plus Environment

13

Environmental Science & Technology

282

Page 14 of 30

Test by comparison: Surface application of consumer product

283 284

To further test the Indoor model, the experimental study of Singer et al. 7 was simulated

285

for a surface application of consumer product in an experimental chamber. A cleaning agent

286

containing 2-butyoxyethanol (2-BE) was sprayed onto a 0.56 m2 section of laminate tabletop in

287

the chamber. After 1 minute, the surface was wiped with paper towels that remained in the room.

288

Total application time was estimated to be 2 minutes; total observation time was 24 hours. The

289

chemical and chamber properties used in the Indoor model to simulate this experiment are given

290

in Tables S-3a and S-3b in the Supporting Information, respectively, and emission data as input

291

to the model are given in Table S-3e of the Supporting Information. For comparison, the

292

ConsExpo model was also used to simulate this cleaning scenario with model inputs as given in

293

Table S-3e of the Supporting Information.

294 295

The human is simulated as remaining in a room for a full hour after using a cleaning

296

product on a hard surface with a 2-minute application period. The Indoor model was used to

297

generate the fugacities in air over the entire time-course.

298 299

Results and Discussion

300 301

Results of model mass balance tests

302 303 304

Mass balance of the Indoor and PBPK models are confirmed in the results shown in Figures 1 and 2, respectively. The time-dependent increase in tissue concentrations and

ACS Paragon Plus Environment

14

Page 15 of 30

Environmental Science & Technology

305

fugacities shows the achievement of apparent steady state after approximately 15 days of

306

constant exposure (Figures S-4b and S-4c in the Supporting Information). It is concluded that the

307

models are mathematically correct and give a complete and consistent mass balance for the

308

chemical in question. The accuracy of the results depend, of course, on the accuracy of the

309

parameter values and on the applicability of the model assumptions. A comparison of model

310

results against experimental observation provides a further test of model accuracy.

311 312

Results of comparison test: Showering

313 314

The volatile organic contaminant, chloroform, entering a shower stall through a domestic

315

water supply was modeled and compared to observations from two sets of experiments 44, 45, 47. A

316

shower water – air transfer efficiency, TE, of 61% (Table S-3a) was calculated for chloroform by

317

the method described in the Supporting Information (Section S-1c-vi). Previous estimates

318

include 56% 45, 61% 48, and 75% of the TE of radon (63%) 34 or 47.25%. To avoid unnecessary

319

repetition, only a small selection of the scenarios from each experiment are modeled.

320 321

In the simulations of the Jo et al experimental showering scenarios 44 (Table 1), the

322

concentration and fugacity of chloroform in the shower stall air increased with time from an

323

assumed pristine state to a maximum when the shower was turned off (Figure S-5a in the

324

Supporting Information). The body burden of chloroform also increased through-out the shower

325

and then declined after water was turned off (Figure S-5a in the Supporting Information). The

326

Indoor model predicted the observed average air concentrations (Table 1); no time dependent air

327

concentrations were available for comparison. At five minutes after showering, the ratio of

ACS Paragon Plus Environment

15

Environmental Science & Technology

Page 16 of 30

328

predicted to measured exhaled air concentration ranged from 7 to 10 (Table 1). The model

329

predictions for exhaled air concentration are dependent upon the breathing rate and the efficiency

330

of uptake through breathing. For example, for the ‘normal’ showering scenario with a water

331

concentration of 0.021 mg/L, decreasing the breathing rate from 1.5 m3/h for the Average Man

332

(Table S-3c in the Supporting Information) to 0.5 m3/h to reflect a lower activity level (a lying

333

man or standing woman, Table S-2e in the Supporting Information) decreased the predicted body

334

burden and hence the concentration in the exhaled air by half (from 40.5 to 19.4 µg/m3). (More

335

detail is given in Figure S-5c in the Supporting Information.) Similarly, decreasing the efficiency

336

of uptake through breathing (50% instead of 70%) decreased the body burden and hence the

337

exhaled air concentration (from 40.5 to 33.4 µg/m3). The combined effect of inhalation and

338

dermal exposure cause an over-prediction of the concentration in exhaled air of less than a factor

339

of four (Table 1); this was because of the mitigating effect of an under-prediction of the effect of

340

dermal exposure (Table 1). Model estimation of the effect of dermal exposure depends upon the

341

selected skin permeability mass transfer coefficient, here assumed to be a constant value of 0.001

342

m/h for all test chemicals. Note that this mass transfer coefficient describes chemical transport

343

into the skin whereas the skin permeability coefficient is routinely measured for transport

344

through the skin 49. Increasing the mass transfer coefficient into skin from the dermal contact

345

medium, and hence the rate of transfer, by a factor of 5 caused the concentration in predicted

346

exhaled air due to dermal exposure to increase from 3.1 to 11.6 µg/m3 which compares favorably

347

with the experimental result of 8.9 µg/m3 (Table 1). Redefining the dermal contact medium as air

348

for the five minutes of drying time had little effect on the overall predicted fate of chloroform for

349

the scenario tested, as discussed in the Supporting Information.

350

ACS Paragon Plus Environment

16

Page 17 of 30

351

Environmental Science & Technology

Observed room air concentrations 47(from Table B2 of Xu 47 and as shown in Figure 1 of

352

Xu and Weisel 45) are well predicted by the Indoor model (Figure 3) for the modeled scenario

353

with subject C. The anomalous reduced room air concentration at 15 minutes was observed by

354

Xu 47 with four of the six subjects. Including this data point in the comparison, the model

355

predictions are within better than a factor of two of observations, i.e., excellent for models of this

356

type. It was assumed that there was no chloroform the air external to the shower. This with the

357

air exchange rate constant may have contributed to the difference between modeled and observed

358

concentrations. The PBPK model shows good agreement with the lowest observed (extracted

359

from Figure 2a of Xu and Weisel 45) exhaled air concentrations (Figure 4). Subject specific

360

exhaled air concentrations were not reported47. This and the absence of a measured breathing rate

361

and efficiency may be the cause of the difference between the predicted and observed values. As

362

shown above in the simulation of the Jo et al 44 experiments and detailed in the Supporting

363

Information (Section S-5), if subject C was breathing more rapidly than estimated, the model

364

would predict higher exhaled air concentrations in closer agreement with observation.

365 366

Results of comparison test: Consumer product applied to a hard surface

367 368

Figure 5 shows the measured and calculated concentrations of 2-BE in the chamber air

369

from both Indoor and ConsExpo models over a 24-hour period. There was an initial rise in

370

concentration of 2-BE followed by a gradual decline. The paper towels used to wipe the surface

371

were an ongoing source after the initial spray and wiping of the tabletop. The Indoor model does

372

not capture the extremely rapid rise in concentration during the first hour but shortly thereafter

373

the simulation is in agreement with the experimental observations 7. This initial under-prediction

ACS Paragon Plus Environment

17

Environmental Science & Technology

Page 18 of 30

374

is approximately a factor of two at the first measured time interval (30 minutes after application).

375

For this scenario, ConsExpo also under-predicts the initial concentrations in the air but over-

376

predicts concentrations thereafter. The Indoor model seemed to capture the exposure better than

377

ConsExpo did especially 2 hours after the application, which may be related to the consideration

378

of chemical partition in different compartments through the fugacity concept in the Indoor

379

model.

380 381

Future considerations

382 383

It has been suggested that the presence of a human occupant may influence the overall

384

fate of chemicals indoors 12. It is expected that the presence of a human occupant will have

385

greatest effect on the overall indoor fate of chemicals with low vapor pressures, consistent with

386

the importance of any sorbent material for increasing chemical residence time in indoor

387

environments, as suggested by Neretnieks et al 50. For the Jo et al 44 experiments, the PBPK

388

model predicted a maximum body burden at 4% of the maximum amount of chloroform in the

389

shower air. In the showering experiments the presence of an occupant appeared to slightly

390

increase the air concentrations of chloroform, although authors considered the difference to be

391

not significant. Additional experimental evidence is needed to clarify the effect of occupancy on

392

the indoor fate of less volatile organic contaminants. It is expected that sorption by humans will

393

be limited by the shorter duration of occupancy relative to stationary sorbing materials. For

394

longer term studies, it may be necessary to consider the effect of the human moving between

395

spaces.

396

ACS Paragon Plus Environment

18

Page 19 of 30

397

Environmental Science & Technology

Advantages

398 399

The Indoor model has the potential, when properly parameterized, to simulate dynamic

400

concentrations indoors. This can be combined with a dynamic model of uptake and distribution

401

within a human to deduce the likely time course of chemical concentrations in tissues and the

402

whole body. This can form the basis of evaluation of both exposure and risk if toxicological data

403

are available. The fugacity formulation allows this to be done seamlessly.

404 405

These models can inform experimental design and protocol. The collection of system

406

dimensions, for example, is often driven by common practice and standard design. Modeling

407

shows the importance of data collection on indoor system dimensions and properties including

408

air exchange rates. Clearly there is a need for further testing of the model against empirical data.

409

With properties specific to the test system model testing can become more rigorous and lead to

410

an improved understanding not possible with assumed, estimated and average values.

411 412 413

Acknowledgement We gratefully acknowledge financial support provided by the Natural Science and

414

Engineering Research Council (NSERC) of Canada as a Collaborative Research and

415

Development (CRD) Grant No 335163-05 to the University of Montreal and Trent University

416

(2008) in collaboration with ExxonMobil Biomedical Sciences Inc (EMBSI) and with a

417

subsequent direct grant from EMBSI to Trent University (2010). The University of Montreal

418

received additional funding from Health Canada, the Institut national de santé publique and

ACS Paragon Plus Environment

19

Environmental Science & Technology

Page 20 of 30

419

Réseau de rech en santé environnementale under the CRD grant. The funds from EMBSI for the

420

CRD grant were provided though its Canadian counterpart, Imperial Oil..

421 422 423

Supporting Information The Indoor and PBPK models are freely available online as stand-alone software from

424

www.trentu.ca/cprg/models. Additional information and results are available free of charge via

425

the Internet at http://pubs.acs.org.

426

ACS Paragon Plus Environment

20

Page 21 of 30

427 428 429 430 431

Environmental Science & Technology

Table 1. The reported average water, shower air and exhaled air concentrations of chloroform from experimental ‘inhalation only’ and ‘normal’ (with both inhalation and dermal sorption) scenarios of Jo et al. 44 selected to test the Indoor and PBPK models and the concentrations predicted by the models.

Water,

Shower stall air, µg/m3

mg/L

Exhaled air, µg/m3, 5 minutes after showering

Reported

Indoor model

average

Reported

PBPK

Predicted/

predicted

model

Reported

average and

predicted

maximum Occupied

0.022

125.9

167; 256

n/r

-

shower stall

0.0356

313.4

270; 415

n/r

-

‘Inhalation only’

0.010

n/r

-

2.4

19.4

8.1

0.021

n/r

-

4.1

40.5

9.9

0.035

n/r

-

8.9

67.4

7.6

0.0053

n/r

-

6.0

11.1

1.8

0.021

n/r

-

13

43.6

3.4

0.036

n/r

-

21

74.7

3.6

13 - 4.1

43.6 - 40.5

0.3

= 8.9

= 3.1

‘Normal’

Implied dermal only 432

0.021

n/r = not reported

433

ACS Paragon Plus Environment

21

Environmental Science & Technology

Page 22 of 30

2284 µg/m3 0.23 mPa

Wallboard

Outdoors 0 µg/m3 0 mPa

17.63 µg/h 371.21 µg/h 17.76 µg/h

0 µg/h

Air 9.3 µg/m3 0.23 mPa

showering flux

8.18 µg/h

Hard Surfaces

8.18 µg/h

2299 µg/m3 0.23 mPa

372.00 µg/h

0.12 µg/h

Room 2

0 nmol/h

4 ×10-3 µg/h 0.64 µg/h 0 nmol/h 0.64 µ/h

Air Circulator

0.64 µg/h

Soft Surfaces 2064 µg/m3 0.23 mPa

6 ×10-4 µg/h

434 435 436 437 438 439

Room 3

Figure 1. Apparent steady state fluxes (µg/h), concentrations (µg/m3), and fugacities (mPa) of the illustrative chemical in the illustrative room after 3 hours of the 5-day showering scenario showing mass balance is achieved. The water fugacity was 10 mPa for a concentration of 1 mg/m3. Degrading reaction losses are shown as dashed arrows.

440

ACS Paragon Plus Environment

22

Page 23 of 30

Environmental Science & Technology

441 ROOM AIR 9.28 µg/m3; 230 µPa

13.920

12.977

4.176 9.744

8.801 ALVEOLAR AIR 208 µPa 8.103

7.159 10-4

10-5

VENOUS BLOOD 18.4 µg/m3; 184 µPa

ARTERIAL BLOOD 20.8 µg/m3; 208 µPa

FAT 1505 µg/m3; 188 µPa

0.367

0.405

0.038 RICHLY PERFUSED 104 µg/m3; 208 µPa

3.483

3.484

0.001 SLOWLY PERFUSED 41.5 µg/m3; 207 µPa

1.739

1.742

0.003

0.297

LIVER 20.9 µg/m3; 29.8 µPa

2.066 0.0027

GIT 2752 µPa

1.772

1.274

0.001

442

SKIN 65.3 µg/m3; 653 µPa 0.130

1.000

DERMAL EXPOSURE MEDIUM 1000 µg/m3; 5025 µPa

0.405

0.0030

0.0003

FOOD 1000 µg/m3; 5025 µPa

443 444 445 446 447 448

Figure 2. Mass balance is shown for the illustrative chemical in the Average Man after 15 days of constant exposure (when steady state conditions appeared to have been achieved) Chemical fluxes (µg/h), concentrations (µg/m3) and fugacities (µPa) shown. Biotransformation reactions shown as dashed arrows. GIT = gastrointestinal tract

ACS Paragon Plus Environment

23

Environmental Science & Technology

Page 24 of 30

350

Air concentration, µg/m3

300 250 200 150 100 50 0 0

5

10

15 20 Time, minutes

25

30

449 Figure 3. The reported (dots) (from Table B2 in Xu 47 ) and Indoor model predicted (solid line) air concentrations of chloroform for subject C during showering.

Exhaled breath concentration, µg/m3

450 451 452

90 80 70 60 50 40 30 20 10 0 0

5

10

15 20 Time, minutes

25

30

453 454 455 456 457

Figure 4. The reported exhaled air concentrations of chloroform for all experimental participants with inhalation-only exposure during showering (extracted from Figure 2a in Xu and Weisel 45) and the exhaled breath concentrations of subject C predicted by PBPK (solid line).

ACS Paragon Plus Environment

24

Page 25 of 30

Environmental Science & Technology

1.6 indoor model ConsExpo measurement

Air Concentration (mg/m3)

1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0

5

10

15

20

25

458

Duration (hrs)

459 460 461 462 463 464 465

Figure 5. Air concentrations in an experimental chamber due to the surface application of 2BE*. Dashed line represents measured, thick solid the Indoor model predictions, and thin solid the ConsExpo predictions. * The measurements are estimations based on Figure 2 in the original study and represent average values for the time periods (0-30 min, 30-60 min, 1-2 hr, 2-4 hr, and 4-24 hr)7

ACS Paragon Plus Environment

25

Environmental Science & Technology

Page 26 of 30

466

References

467 468

(1) Bennett, D. H.; Furtaw, E. J. Fugacity-based indoor residential pesticide fate model. Environ. Sci. Technol. 2004, 38 2142-2152.

469 470 471

(2) Colombo, A.; De Bortoli, M.; Knöppel, H.; Pecchio, E.; Vissers, H. Adsorption of selected volatile organic compounds on a carpet, a wall coating, and a gypsum board in a test chamber. Indoor Air 1993, 3 276-282.

472 473

(3) Mackay, D.; Paterson, S. Fugacity models of indoor exposure to volatile chemicals. Chemosphere 1983, 12 143-154.

474 475 476

(4) Matoba, Y.; Yoshimura, J.; Ohnishi, J.; Mikami, N.; Matsuo, M. Development of the simulation model InPest for prediction of the indoor behavior of pesticides. J. Air Waste Manage. Assoc. 1998, 48 969-978.

477 478

(5) Nazaroff, W. W.; Cass, G. R. Mathematical modeling of chemically reactive pollutants in indoor air. Environ. Sci. Techol. 1986, 20 (9), 924-934.

479 480 481

(6) Robson, M.; Melymuk, L.; Csiszar, S. A.; Giang, A.; Diamond, M. L.; Helm, P. A. Continuing sources of PCBs: The significance of building sealants. Environ. Internat. 2010, 36 506-513.

482 483 484

(7) Singer, B. C.; Destaillats, H.; Hodgson, A. T.; Nazaroff, W. W. Cleaning products and air fresheners: Emissions and resulting concentrations of glycol ethers and terpenoids. Indoor Air 2006, 16 179-191.

485 486 487

(8) Shin, H.-M.; McKone, T. E.; Tulve, N. S.; Clifton, M. S.; Bennett, D. H. Indoor residence times of semivolative organic compounds: Model estimation and field evaluation. Environ. Sci. Technol. 2013, 47 (2), 859-867.

488 489

(9) Weschler, C. J. Changes in indoor pollutants since the 1950s. Atmos. Environ. 2009, 43 153-169.

490 491

(10) Weschler, C. J.; Nazaroff, W. W. Semivolatile organic compounds in indoor environments. Atmos. Environ. 2008, 42 9018-9040.

492 493

(11) Wolkoff, P. Indoor air pollutants in office environments: Assessment of comfort, health, and performance. Internat. J. Hygiene Environ. Health 2013, 216 371-394.

494 495 496

(12) Zhang, X.; Arnot, J. A.; Wania, F. Model for screening-level assessment of near-field human exposure to neutral organic chemicals released indoors. Environ. Sci. Technol. 2014, 48 (20), 12312-12319; 10.1021/es502718k

ACS Paragon Plus Environment

26

Page 27 of 30

Environmental Science & Technology

497 498 499

(13) Zhang, X.; Diamond, M. L.; Ibarra, C.; Harrad, S. Multimedia modeling of polybrominated diphenyl ether emissions and fate indoors. Environ. Sci. Technol. 2009, 43 28452850.

500 501 502 503

(14) Brandsma, S. H.; de Boer, J.; van Velzen, M. J. M.; Leonards, P. E. G. Organophosphorus flame retardants (PFRs) and plasticizers in house and car dust and the influence of electronic equipment. Chemosphere 2014, 116 3-9; 10.1016/j.chemosphere.2014.02.036

504 505 506 507

(15) Fang, M.; Webster, T. F.; Gooden, D.; Cooper, E. M.; McClean, M. D.; Carignan, C.; Makey, C.; Stapleton, H. M. Investigating a novel flame retardant known as V6: Measurements in baby products, house dust, and car dust. Environ. Sci. Technol. 2013, 47 (9), 4449-4454; 10.1021/es400032v

508 509 510

(16) Yoshida, T.; Matsunaga, I. A case study on identification of airborne organic compounds and time courses of their concentrations in the cabin of a new car for private use. Environ. Internat. 2006, 32 (1), 58-79; DOI 10.1016/j.envint.2005.04.009

511 512 513

(17) Allen, J. G.; Stapleton, H. M.; Vallarino, J.; McNeely, E.; McClean, M. D.; Harrad, S. J.; Rauert, C. B.; Spengler, J. D. Exposure to flame retardant chemicals on commercial airplanes. Environ. Health 2013, 12 10.1186/1476-069x-12-17

514 515 516

(18) Allen, J. G.; Sumner, A. L.; Nishioka, M. G.; Vallarino, J.; Turner, D. J.; Saltman, H. K.; Spengler, J. D. Air concentrations of PBDEs on in-flight airplanes and assessment of flight crew inhalation exposure. J. Expo. Sci. Environ. Epidem. 2013, 23 (4), 337-342; 10.1038/jes.2012.62

517 518 519

(19) Strid, A.; Smedje, G.; Athanassiadis, I.; Lindgren, T.; Lundgren, H.; Jakobsson, K.; Bergman, A. Brominated flame retardant exposure of aircraft personnel. Chemosphere 2014, 116 83-90; 10.1016/j.chemosphere.2014.03.073

520 521 522 523

(20) Wisthaler, A.; Strøm-Tejsen, P.; Fang, L.; Arnaud, T. J.; Hansel, A.; Tilmann D. Märk; Wyon, D. P. PTR-MS assessment of photocatalytic and sorption-based purification of recirculated cabin air during simulated 7-h flights with high passenger density. Environ. Sci. Technol. 2007, 41 (1), 229-234; DOI 10.1021/es060424e

524 525 526

(21) Delmaar, J. E.; Park, M. V. D. Z.; van Engelen, J. G. M. ConsExpo 4.0, Consumer Exposure and Uptake Models Program Manual 2005 RIVM report 320104004 National Institute for Public Health and the Environment of the Netherlands (RIVM)

527 528

(22) Guo, Z. Review of indoor emission source models. Part 1. Overview. Indoor Air 2002, 120 533-549.

529 530 531

(23) Liagkourdis, I.; Cousins, I. T.; Palm Cousins, A. Emissions and fate of brominated flame retardants in the indoor environment: A critical review of modelling approaches. Sci. Total. Environ. 2014, 491-492 87-99; 10.1016/j.scitotenv.2014.02.005

ACS Paragon Plus Environment

27

Environmental Science & Technology

Page 28 of 30

532 533 534

(24) OECD. Description of Existing Models and tools used for exposure assessment: Results of an OECD Survey 2012 Series of Testing and Assessment Report No. 182 ENV/JM/MONO(2012)37

535 536 537

(25) Park, M. V. D. Z.; Delmaar, J. E.; van Engelen, J. G. M. Comparison of consumer exposure modelling tools 2006 RIVM report 320104006 National Institute for Public Health and the Environment of the Netherlands (RIVM)

538 539

(26) Matoba, Y.; Ohnishi, J.; Matsuo, M. Indoor simulation of insecticides supplied with an electric vaporizer by the fugacity model. Chemosphere 1994, 28 767-786.

540 541

(27) Matoba, Y.; Ohnishi, J.; Matsuo, M. Temperature- and humidity-dependency of pesticide behavior in indoor simulation. Chemosphere 1995, 30 933-952.

542 543

(28) Nazaroff, W. W. Inhalation intake fraction of pollutants from episodic indoor emissions. Building and Environment 2008, 43 269-277.

544 545 546

(29) Himmelstein, K. J.; Lutz, R. J. Review of the application of physiologically based pharmacokinetic modeling. J. Pharmacokinet. Biopharm. 1979, 7 (2), 127-145; 10.1007/bf01059734

547 548

(30) Rowland, M. Physiologic pharmacokinetic models - Relevance, experience, and future trends. Drug Metabolism Reviews 1984, 15 (1-2), 55-74; 10.3109/03602538409015057

549 550 551

(31) Bjorkman, S. Reduction and lumping of physiologically based pharmacokinetic models: Prediction of the disposition of fentanyl and pethidine in humans by successively simplified models. J. Pharmacokinet. Pharmacodyn. 2003, 30 (4), 285-307; 10.1023/a:1026194618660

552 553 554

(32) Bjorkman, S. Prediction of the volume of distribution of a drug: which tissue-plasma partition coefficients are needed? J. Pharm. Pharmacol. 2002, 54 (9), 1237-1245; 10.1211/002235702320402080

555 556 557

(33) Lipscomb, J. C.; Haddad, S.; Poet, T.; Krishnan, K. Physiologically-based pharmacokinetic (PBPK) models in toxicity testing and risk assessment in New Technologies for Toxicity Testing; Balls, M.; Combes, R. D.; Bhogal, N.; 2012; pp 76-95.

558 559

(34) McKone, T. E. Human exposure to volatile organic compounds in household tap water: The indoor inhalation pathway. Environ. Sci. Technol. 1987, 21 (12), 1194-1201.

560 561 562

(35) Shin, H. M.; McKone, T. E.; Bennett, D. H. Intake fraction for the indoor environment: A tool for prioritizing indoor chemical sources. Environ. Sci. Technol. 2012, 46 (18), 10063-10072; 10.1021/es3018286

563 564 565

(36) Wenger, Y.; Li, D.; Jolliet, O. Indoor intake fraction considering surface sorption of air organic compounds for life cycle assessment. Internat. J. Life Cycle Assess. 2012, 17 (7), 919931; 10.1007/s11367-012-0420-0

ACS Paragon Plus Environment

28

Page 29 of 30

Environmental Science & Technology

566 567

(37) Isaacs, K.; Glen, G.; D., S.; Bennett, D. Development and evaluation of a fugacity-based source-to-concentration module for SHEDS-Multimedia. Poster WP-41 2012.

568 569 570

(38) Haddad, S.; Tardif, R.; Boyd, J.; Krishnan, K. Physiologically based modeling of pharmacokinetic interactions in chemical mixtures. Ch 4 in Quantitative Modeling in Toxicology; Krishnan, K.; Andersen, M.; John Wiley & Sons: 2010; pp 83-105.

571 572

(39) Mackay, D. Multimedia environmental models: The fugacity approach. Lewis Publishers: Boca Raton, FL; 2001.

573 574 575

(40) Poulin, P.; Krishnan, K. An algorithm for predicting tissue-blood partition coefficients of organic chemicals from n-octanol-water partition coefficient data. J. Toxicol. Environ. Health 1995, 46 (1), 117-129.

576 577 578

(41) Buist, H. E.; de Wit-Bos, L.; Bouwman, T.; Vaes, W. H. J. Predicting blood:air partition coefficients using basic physicochemical properties. Reg. Toxicol. Pharmacol. 2012, 62 (1), 2328; 10.1016/j.yrtph.2011.11.019

579 580

(42) Paterson, S.; Mackay, D. A steady state fugacity based pharmacokinetic model with simultaneous multiple exposure routes. Environ. Toxicol. Chem. 1987, 6 395-408.

581 582 583

(43) Cahill, T. M.; Cousins, I.; MacKay, D. Development and application of a generalized physiologically based pharmacokinetic model for multiple environmental contaminants. Environ. Toxicol. Chem. 2003, 22 (1), 26-34; 10.1897/1551-5028(2003)022

584 585

(44) Jo, W. K.; Weisel, C. P.; Lioy, P. J. Routes of chloroform exposure and body burden from showering with chlorinated tap water. Risk Anal. 1990, 10 (4), 575-580.

586 587

(45) Xu, X.; Weisel, C. P. Human respiratory uptake of chloroform and haloketones during showering. J. Expo. Anal. Env. Epid. 2005, 15 (1), 6-16; 10.1038/sj.jea.7500374

588 589

(46) Xu, X.; Weisel, C. P. Inhalation exposure to haloacetic acids and haloketones during showering. Environ. Sci. Technol. 2003, 37 (3), 569-576; 10.1021/es025747z

590 591

(47) Xu, X. Dermal and inhalation exposure to disinfection by-products in “drinking water”. Ph.D. Dissertation, Rutgers University New Brunswick, NJ. 2002.

592 593 594

(48) Chinery, R. L.; Kevin, G. A. A compartmental model for the prediction of breath concentration and absorbed dose of chloroform after exposure while showering. Risk Anal. 1993, 13 (1), 51-62.

595 596 597

(49) Abdallah, M. A.-E.; Pawar, G.; Harrad, S. Evaluation of 3D-human skin equivalents for assessment of human dermal absorption of some brominated flame retardants. Environ. Internat. 2015, 84 64-70; 10.1016/j.envint.2015.07.015

ACS Paragon Plus Environment

29

Environmental Science & Technology

598 599 600

Page 30 of 30

(50) Neretnieks, I.; Christiansson, J.; Romero, L.; Dagerholt, L.; Yu, J.-W. Modelling of emission and re-emission of volatile organic compounds from building materials with indoor air applications. Indoor Air 1993, 3 (1), 2-11; 10.1111/j.1600-0668.1993.t01-3-00002.x

601

602

For Table of Contents Only

603 Indoor Model Wallboard Outdoors

Air

Hard Surfaces

Room 2

Air Circulator

Soft Surfaces

PBPK Model

Room 3

604 605 606

ACS Paragon Plus Environment

30