IPCS INCHEM Home

WHO FOOD ADDITIVES SERIES: 48

SAFETY EVALUATION OF CERTAIN
FOOD ADDITIVES AND CONTAMINANTS

POLYCHLORINATED DIBENZODIOXINS,
POLYCHLORINATED DIBENZOFURANS, AND
COPLANAR POLYCHLORINATED BIPHENYLS

First draft prepared by R. Canady1, K. Crump2, M. Feeley3, J. Freijer4, M. Kogevinas5, R. Malisch6, P. Verger7, J. Wilson8 and M. Zeilmaker9
1
Office of Plant & Dairy Foods and Beverages, Center for Food Safety & Applied Nutrition, Food and Drug Administration, Washington, DC, USA
2Ruston, LA, USA
3Bureau of Chemical Safety, Food Directorate, Health Products and Food Branch, Health Canada, Ottawa, Ontario, Canada
4National Institute of Public Health and the Environment, Bilthoven, Netherlands;
5Respiratory and Environmental Health Research Unit, Municipal Institute of Medical Research, Barcelona, Spain
6Chemisches und Veterinäruntersuchungsamt, Freiburg, Germany
7Scientific Directorate on Human Nutrition and Food Safety, National Institute for Agricultural Research, Paris, France
8Center for Risk Management, Resources for the Future, Washington DC, USA
9Center for Substances and Risk Assessment, National Institute of Public Health and the Environment, Bilthoven, Netherlands

Explanation

Biological data

Biochemical aspects

Absorption, biotransformation and excretion

Absorption

Uptake and distribution in the body

Elimination

Transport across the placenta

Physiologically based pharmacokinetics modelling

Biochemical effects

Induction of CYP 1A1 and CYP 1A2

Induction of CYP 1B1

Induction of UGT1, PA12 and TGF-alpha?

TGF-alpha and EGF pathway

Thyroid hormones

Retinoids

Hepatic porphyria

Toxic equivalents factors

Toxicological studies

Acute toxicity

Short-term studies of toxicity

Long-term studies of toxicity and carcinogenicity

Genotoxicity

Reproductive toxicity

Multigeneration studies

Single-generation studies

Developmental toxicity

Special studies

Neurotoxicity

Immunotoxicity

Effects on the thyroid

Vitamin A status

Observations in humans

Cancer

Cancer incidence and mortality rates in the population of Seveso

Exposure in chemical plants

Exposure of the general population and commercial pesticide and herbicide applicators

Cancer in children

Conclusions

Effects other than cancer

Reproductive effects

Developmental effects

Chloracne

Gastroinestinal effects and effects on hepatic enzymes

Effects on thyroid function

Diabetes

Cardiovascular effects and lipid concentrations

Neuropsychological effects

Conclusions

Analytical methods

Chemistry

Description of methods

Congeners to be determined

Availability of official methods

Upper- and lower-bound concentrations

Required sensitivity and analytical approaches

Conclusions and recommendations

Maximum limits, action levels, target levels

General requirements for monitoring adherence to maximum levels

General requirements for monitoring background contamination

Special requirements for GC–MS methods

Harmonized quality criteria for cell-based and kit-based bioassays

Combination of bioassays and GC–MS analysis

Sampling protocols

Personnel

Representative sample

Packaging, transport and storage of aggregate and laboratory samples

Human milk samples

Sealing and labelling

Edible parts

Effects of processing

Carry-over from packaging and filter material

Smoking, frying and broiling

Levels and patterns of contamination of food commodities

Results of surveys

Distribution in foods

Estimated dietary intake

Introduction and background

Calculations of intake

Relevant period of intake

Methods used

Intake calculations

Compounds

Distributions of concentrations

National diets

International diets

Hypothetical maximum limits

National estimates of intake

Central trend

High intakes

Dietary sources

Regional estimates of intake

Central trend

High intakes

Dietary sources

Effects of regulatory limits

Dioxin concentrations

Dioxin intakes

Biomonitoring

Correlations between different matrices

WHO-coordinated study of concentrations in human milk: Comparison of countries

WHO-coordinated study: Time trends

Variation in dioxin content of breast milk samples

Dependence of toxic equivalents on age

Data from countries not participating in WHO field studies

Pattern of congeners

Variation in body burden as an estimate of variation in intake

Summary

Prevention and control

Preharvest control

Decontamination

Dose–response analysis and estimation of risk

Safety assessment

Pivotal toxicological data

Adjustment from bolus dose to equivalent body burden in studies of developmental toxicity

Selection of critical studies

Background body burdens in laboratory animals

Selection of safety factors

Provisional tolerable monthly intake

Dose–response relationship above the tolerable intake

Effects of uncertainty in estimates of intake

Quantitative uncertainty in human cancer risk assessment

Short-term effects other than cancer

Estimate of lowest effect level for developmental effects from data on humans

Dose–response relationships above the tolerable intake based on data for experimental animals

'Benchmark' dose modelling

Comments

Evaluation

References

1. EXPLANATION

1.1 Introduction

Polychlorinated dibenzodioxins (PCDDs) and polychlorinated dibenzofurans (PCDFs) are by-products of combustion and of various industrial processes, and they are widely present in the environment. Polychlorinated biphenyls (PCBs) were manufactured in the past for a variety of industrial uses, notably as electrical insulators or dielectric fluids and specialized hydraulic fluids. Most countries banned manufacture and use of PCBs in the 1970s; however, past improper handling of PCBs constitutes a continuing source of these compounds in the environment, and disposal of equipment now in use poses some risk of further contamination.

Neither PCDDs nor PCDFs have been evaluated previously by the Committee. PCBs were evaluated at the thirty-fifth meeting, when a provisional tolerable weekly intake (PTWI) could not be established because of the limitations of the available data and the ill-defined nature of the materials that were used in feeding studies (Annex 1, reference 88).

PCDDs, PCDFs and coplanar PCBs were evaluated at the present meeting on the basis of a request by the Codex Committee on Food Additives and Contaminants (CCFAC) to evaluate the risks associated with their presence in food.

The Committee evaluated the PCDDs, PCDFs and coplanar PCBs for which toxic equivalency factors (TEFs) for mammals have been derived by WHO (Ahlborg et al., 1994). Table 1 lists the compounds that were considered and their assigned TEF values. The TEF approach relates the toxicity of all chemicals in the series to that of 2,3,7,8-tetrachlorinated dibenzodioxin (TCDD), one of the most potent of the chemicals on which most toxicological and epidemiological information was available. Use of the TEF concept rests on the assumption that PCDDs, PCDFs and coplanar PCBs have a common mechanism of action, which involves binding to the aryl hydrocarbon (Ah) receptor, an intracellular receptor protein. This binding is considered to be the necessary first, but not sufficient, step in expressing the toxicity of these compounds. Many uncertainties exist in use of the TEF approach for human risk assessment, but pragmatically it is the most feasible approach available.

Table 1. Compounds considered and the toxic equivalency factor assigned by WHO

Compound

Abbreviation

Toxic equivalency factor

Polychlorinated dibenzodioxins

 

 

2,3,7,8-Tetrachlorodibenzodioxin

TCDD

1

1,2,3,7,8-Pentachlorodibenzodioxin

1,2,3,7,8-PeCDD

1

1,2,3,4,7,8-Hexachlorodibenzodioxin

1,2,3,4,7,8-HxCDD

0.1

1,2,3,6,7,8-Hexachlorodibenzodioxin

1,2,3,6,7,8-HxCDD

0.1

1,2,3,6,7,9-Hexachlorodibenzodioxin

1,2,3,6,7,9-HxCDD

0.1

1,2,3,4,6,7,8-Heptachlorodibenzodioxin

1,2,3,4,6,7,8-HpCDD

0.01

Octachlorodibenzodioxin

OCDD

0.0001

Polychlorinated dibenzofurans

 

 

2,3,7,8-Tetrachlorodibenzofuran

2,3,7,8-TCDF

0.1

1,2,3,7,8-Pentachlorodibenzofuran

1,2,3,7,8-PeCDF

0.05

2,3,4,7,8-Pentachlorodibenzofuran

2,3,4,7,8-PeCDF

0.5

1,2,3,4,7,8-Hexachlorodibenzofuran

1,2,3,4,7,8-HxCDF

0.1

1,2,3,6,7,8-Hexachlorodibenzofuran

1,2,3,6,7,8-HxCDF

0.1

1,2,3,7,8,9-Hexachlorodibenzofuran

1,2,3,7,8,9-HxCDF

0.1

2,3,4,6,7,8-Hexachlorodibenzofuran

2,3,4,6,7,8-HxCDF

0.1

1,2,3,4,6,7,8-Heptachlorodibenzofuran

1,2,3,4,6,7,8-HpCDF

0.01

1,2,3,4,7,8,9-Heptachlorodibenzofuran

1,2,3,4,7,8,9-HpCDF

0.01

Octochlorodibenzofuran

OCDF

0.0001

‘Non-ortho’ polychlorinated biphenyls

 

 

3,3΄,4,4΄-Tetrachlorobiphenyl (polychlorinated biphenyl #77)

3,3΄,4,4΄-TCB

0.0001

3,4,4΄,5,-Tetrachlorobiphenyl (polychlorinated biphenyl #81)

3,4,4΄,5-TCB

0.0001

3,3΄,4,4΄,5-Pentachlorobiphenyl (polychlorinated biphenyl #126)

3,3΄,4,4΄,5-PeCB

0.1

3,3΄,4,4΄,5,5΄-Hexachlorobiphenyl (polychlorinated biphenyl #169)

3,3΄,4,4΄,5,5΄-HxCB

0.01

‘Mono-ortho’ polychlorinated biphenyls

 

 

2,3,3΄,4,4΄-Pentachlorobiphenyl (polychlorinated biphenyl #105)

2,3,3΄,4,4΄-PeCB

0.0001

2,3,4,4΄,5-Pentachlorobiphenyl (polychlorinated biphenyl #114)

2,3,4,4΄,5-PeCB

0.0005

2,3΄,4,4΄,5-Pentachlorobiphenyl (polychlorinated biphenyl #118)

2,3΄,4,4΄,5-PeCB

0.0001

2,3΄,4,4΄,5’-Pentachlorobiphenyl (polychlorinated biphenyl #123)

2,3΄,4,4΄,5΄-PeCB

0.0001

2,3,3΄,4,4΄,5-Hexachlorobiphenyl (polychlorinated biphenyl #156)

2,3,3΄,4,4΄,5-HxCB

0.0005

2,3,3΄,4,4΄,5΄-Hexachlorobiphenyl (polychlorinated biphenyl #157)

2,3,3΄,4,4΄,5΄-HxCB

0.0005

2,3΄,4,4΄,5,5΄-Hexachlorobiphenyl (polychlorinated biphenyl #167)

2,3΄,4,4΄,5,5΄-HxCB

0.00001

2,3,3΄,4,4΄,5,5΄-Heptachlorobiphenyl (polychlorinated biphenyl #189)

2,3,3΄,4,4΄,5,5΄-HpCB

0.00001

Two documents were particularly important in this evaluation. A WHO consultation held in 1998 (van Leeuwen & Younes, 2000) established a tolerable daily intake (TDI) of 1–4 pg/kg bw, which was applied to the toxic equivalents of PCDDs, PCDFs and coplanar PCBs. The TDI was based on the results of a number of studies of developmental toxicity, in which pregnant rats were given TCDD by gavage, and immunological toxicity. The present Committee used this assessment as the starting point for its evaluation, taking into account newer studies that provided information on:

The second is a position paper on dioxins, developed for the CCFAC at its thiry-third session (Codex Alimentarius, 2001), which summarizes levels of exposure and values derived in safety assessments and explores the arguments for and against setting maximum limits. In addition, comprehensive evaluations have been conducted by several organizations, including IARC (1997), the Agency for Toxic Substances and Disease Registry (1998) in the USA, the European Union (1999, 2000a,b,c) and the Environmental Protection Agency (2000a) in the USA.

1.2 Compounds considered

The 29 compounds listed in Table 1 are covered by the assessment. These compounds have similar resistance to environmental and metabolic degradation and solubility in body fat, and they share a unique spectrum of toxic responses initiated by interaction with the Ah receptor found in many tissues in the body.

PCDDs and PCDFs are by-products of combustion and of various industrial processes, and they are widely present in the environment. The subset of these compounds considered in this assessment comprises those with chlorine substitutions at the 2, 3, 7 and 8 positions. The prototypical member of this group, TCDD, is generally regarded as one of the most potent toxins known. 1,2,3,7,8-Pentachlorodibenzodioxin is of a similar potency, while the other members of the subset are 10–10 000 times less toxic.

The 12 PCBs included in this assessment are considered to share dioxin-like properties and have either one or no chlorine substitutions in the ortho positions. Non-ortho- and mono-ortho-substituted PCBs in the environment and in foods generally comprise a small percentage of the total PCB contamination. The dioxin-like toxicity of these 12 PCBs is 10–100 000 less than that of TCDD.

1.3 Toxic equivalents

The assumption made throughout this document is that the 29 compounds have a common mechanism of action and all the compounds act through this mechanism. This assumption allows consideration of a broader range of data on toxicity, particularly in the case of human poisoning incidents. The larger benefit of the assumption is that it allows data on exposure to the 29 compounds to be summarized in a single description. In the absence of this assumption, individual effects, potency, and the sufficiency of data would have to be considered for each compound. Use of toxic equivalents to broaden the database on toxicity and to simplify the descriptions of risk comes, however, at the cost of reducing the possibility of projecting the uncertainty in the evaluations of toxicity and risk into the risk characterization. For example, use of data on the effects of exposure to furans alone to estimate risk relies on the validity and accuracy of the TEF for the furans (see section 2.1.4), but the uncertainty of the TEFs for individual furans is not explicitly taken into consideration.

The accuracy of estimates of toxic equivalents is uncertain for similar reasons and for the additional reason that other compounds present in the environment may affect the biological response through the assumed common mechanism. Thus, compounds such as brominated and chlorobrominated analogues of PCDDs, PCDFs, naphthalenes, diphenyl ethers, diphenyl toluenes, phenoxyanisoles, biphenyl anisoles, xanthenes, xanthones, anthracenes, fluorenes, dihydroanthracenes, biphenyl methanes, phenylxylylethanes, dibenzothiophenes, quaterphenyls, quater-phenyl ethers and biphenylenes could all affect the ‘true’ toxic equivalents of food.

Furthermore, toxic equivalence is assumed to be simply additive in all cases, despite evidence that the effects of some compounds in environmental mixtures are less than additive, greater than additive (synergistic), or antagonistic (reduce the adverse biological response). There is essentially no information on the accuracy of estimates of toxic equivalents in predicting the true adverse biological response to the various mixtures found in food, nor is there an adequate basis for estimating the uncertainty of the estimate.

Most of the evidence for the toxicity of the 29 compounds comes from studies of TCDD. The toxic equivalents method is based on the assumption that the toxicity of TCDD is equal to or greater than that of any of the other 28 compounds. In the TEF scheme, a dose of TCDD of 1 pg/kg bw is considered to be equivalent to a toxic equivalence of 1 pg/kg bw. The contribution of TCDD to the estimated toxic equivalents of a food is typically less than 10%. Nonetheless, the Committee used the toxic equivalents method to allow inclusion of data such as that from the Yusho and Yu-cheng incidents of rice oil poisoning (see section 2.3.2), for which the toxic equivalence was due entirely to furans and PCBs. Toxic equivalence was also used to describe intake from food and as a basis for estimating tolerable intake.

1.4 General considerations on exposure to dioxins

Persons can be expossed to PCDDs, PCDFs and coplanar PCBs occupationally, accidentally, or in the environmental (background). Exposure to background contamination can occur by inhalation, ingestion, or contact with contaminated soil. Assessments of exposure by the European Commission (2000a) and the EPA (2000a) in the USA showed that > 90% of the exposure of a typical person to PCDDs, PCDFs and coplanar PCBs came from food and predominantly from animal fat (Bund/Länder Arbeitsgruppe Dioxine, 1993; European Union, 1999; Environmental Protection Agency, 2000a; European Union, 2000a,b; van Leeuwen & Younes, 2000). The contamination of animal fat is thought to be derived largely from feed (rather than, for example, soil contact or inhalation of air by food animals), and therefore animal feed is a potential control point for reduction of the intake of PCDDs, PCDFs and coplanar PCBs from the food chain (European Union, 2000c).

1.4.1 Sources of PCDDs, PCDFs and coplanar PCBs that contaminate food

Dioxins and furans are released into the air during combustion processes such as industrial and municipal waste incineration (including burning of household waste in some areas), metal recycling and refining (smelting) and burning of fuels like wood, coal, gasoline, or oil. Dioxins and furans can also be formed from natural sources (for example, during forest fires). Chlorine bleaching of pulp and paper, certain types of chemical manufacture and processing and other industrial processes all can create small quantities of dioxins and furans.

The sources of PCBs are different from those of PCDDs and PCDFs, in that there was substantial commercial production of PCBs. PCBs have been released to the environment over the past 70 years from PCB-containing equipment in industrial discharges and by improper use and disposal of equipment containing PCBs. Because their manufacture and use has been banned in most countries, the predominant source of PCBs now is the environmental reservoir from past releases.

Federal governments, industry and environmental interest groups have worked together for over a decade to reduce emissions of PCDDs, PCDFs and coplanar PCBs. However, because these compounds are extremely persistent, past releases remain in the environment as contaminated soils and sediments and will take decades to decline. The contribution of these ‘environmental reservoirs’ to food contamination has not been quantified; however, on the basis of the volumes of past release and the persistence in the environment, environmental reservoirs will become the single largest source of these compounds to food, as industrial and waste-stream emmissions are reduced. As the environment is in some sense the ‘proximal’ source of many if not most foods, both proximal and release sources should be considered in efforts to find the most effective means for reducing exposure.

The relative contribution of PCDDs, PCDFs and coplanar PCBs to the total environmental load from various sources has changed substantially over the past decades. Furthermore, the relative importance of sources varies from one country to another. In the past and in industrialized countries, the chemical industry was the main source of releases of PCDDs and PCDFs into the environment. Today, the main (quantified) releases are from combustion processes.

UNEP (1999) has started to collect data from national and regional inventories of dioxin. It became evident that there were no harmonized methods for establishing inventories. As the Stockholm Convention on POPs will require continuous minimization of releases of these compounds, UNEP (2001) has offered a standardized toolkit for establishing inventories of PCDDs, PCDFs and coplanar PCBs. Most inventories cover emissions to the air only; less information is available on releases of residues and products to land and water. Most of the information comes from the Northern Hemisphere, and the sources in developing countries have not been quantified. Changes in techniques for waste incineration have reduced exposure in industrialized countries, but the role of reservoirs remains to be evaluated. Iron and steel manufacture is an important contributor in many countries, but not all industrialized countries include this important sector in their inventories.

1.4.2 Pathways to food

Transfer of environmental contamination into animal feed commonly results in the appearance of PCDDs, PCDFs and coplanar PCBs as contaminants in fat-containing animal products, meat and milk. Feed, food-producing animals and food products may become contaminated in various ways, including deposition of emissions from various sources on farmland, burning of contaminated raw material for direct drying, blending of feedstuffs with contaminated products, application of contaminated pesticides, detergents, or disinfectants, contact with wooden materials treated with wood preservatives, application of sewage sludge to fields, flooding of pastures, contamination of water with wastewater and effluents, food processing, or migration from chlorine-bleached packaging material.

(a) Environmental contamination

Widespread environmental contamination with PCDDs, PCDFs and coplanar PCBs remains after past releases. As the half-lives of some of these compounds in the environment are decades or longer, the environmental contamination is likely to persist for some time. As a result, most of the contamination of food by PCDDs, PCDFs and coplanar PCBs is due to their occurrence in the environment and is not easily traced to the original source.

Food may become contamined via many pathways, including direct deposition from the air onto leafy plants used in feed and ingestion of contaminated soil by herbivores (e.g. the roots of grass pulled during grazing). In general, PCDDs, PCDFs and coplanar PCBs do not accumulate in plant matter other than by external deposition from the air; for example, most plants do not take up PCDDs, PCDFs and coplanar PCBs from the soil but can carry them on their surfaces to differing degrees. Potatoes and carrots can take up these compounds from contaminated soil into their outer layers. The only plants for which a mechanism for uptake and distribution has been demonstrated are courgette and pumpkin. Feed may also be contaminated (European Union, 2000c). Owing to the ubiquity of contamination with PCDDs, PCDFs and coplanar PCBs and the low limits of detection required to identify biologically relevant concentrations, there is substantial uncertainty about the predominant pathways by which these compounds enter the food supply.

(b) Accidents

During the past few decades, heavy exposure to dioxins and furans has occurred in isolated incidents of contamination or release. Well-studied examples of environmental releases include the exposure of the local population at Seveso, Italy (Pocchiari et al., 1979; Bertazzi & di Domenico, 1994), and from fires in PCB-filled electrical equipment, such as in the Binghamton State Office Building in New York State, USA (Fitzgerald et al., 1986, 1989). Heavy exposure, with toxic effects, has also been caused by contaminated foods. Known examples are the contamination of edible oils, such as in the Yusho (Japan) and Yu-cheng (Taiwan) food poisoning episodes (Rogan et al., 1988; Kuratsune et al., 1996; see section 2.3.2), which involved exposure to concentrations of dioxin or furan at least three to four orders of magnitude higher than the highest normally found in foods.

Incidents of lighter contamination, with no known toxic effects, have been reported, which include ingestion of a naturally contaminated feed additive (a form of clay) which led to elevated concentrations of dioxin in catfish and poultry (Rappe et al., 1998; Ferrario et al., 1999; Holcomb et al., 1999; Eljarrat et al., 2000; Jobst & Aldag, 2000; Malisch, 2000a); ingestion of a feed additive heavily contaminated with PCB waste that led to contaminated poultry, eggs, milk and meat in Belgium (Broeckaert & Bernard, 2000; Belgian Federal Government, 2001); and three incidents of agricultural practices that led to contamination of animal feeds and food: contamination of citrus pulp pellets as a result of use of heavily contaminated lime for neutralization (Malisch et al., 1999; Traag et al., 1999; Malisch, 2000b; Malisch et al., 2000), contamination of grass meal as a result of use of contaminated wood for direct drying (European Union, 2000c) and contamination of choline chloride as result of use of contaminated wood as a carrier (European Union, 2000c).

These cases show that food can become contaminated in a variety of ways. After the successful reduction of emissions of PCDDs, PCDFs and coplanar PCBs into the environment in the 1970s, 1980s and 1990s, attention must now be focused on animal feed and the pathways to feed in order to reduce the amounts of these compounds entering the food supply.

2. BIOLOGICAL DATA

2.1 Biochemical aspects

2.1.1 Absorption, distribution, biotransformation and excretion

(a) Absorption

As fat is efficiently absorbed from the gastrointestinal tract, dioxin-like compounds administered in a fatty matrix can be expected to pass easily into the blood. Experiments in rats showed approximately 90% absorption of 2,3,7,8-TCDF after oral administration of a single dose in a 1:1 ethanol:vegetable oil mixture (Birnbaum et al., 1980) and 70–85% absorption of 2,3,4,7,8-PeCDF (Yoshimura et al., 1986; Brewster & Birnbaum, 1987; Kanimura et al., 1988). Similarly, (mean) absorption fractions of 0.84 (range, 0.66–0.93) after oral administration in corn oil (Rose et al., 1976)) and 0.88 (standard deviation, 1.7) after oral administration in a 1:1:3 solution of vegetable oil, ethanol and water (Diliberto et al., 1996) have been reported for TCDD in rats.

In contrast to TCDD and 2,3,7,8-TCDF, OCDD is poorly absorbed, 2–15% of a single dose being absorbed after administration by gavage in a 1:1 ortho-dichlorobenzene:corn oil mixture (Birnbaum & Couture, 1988; Couture et al., 1988). Furthermore, the absorption of a single oral dose of 1,2,3,7,8-PeCDD was found to vary from 19 to 71% (Wacker et al., 1986).

Little may be absorbed from more complex matrices such as the diet. As little as 50–60% of a dose of TCDD in the diet was absorbed (Fries & Marrow, 1975).

In a study in which TCDD was given orally in corn oil to a volunteer, > 87 % was absorbed (Poiger & Schlatter, 1986). This figure is comparable with the near complete absorption of dioxins, furans and PCBs by nursing infants from mother’s milk (McLachlan, 1993; Dahl et al., 1995).

(b) Uptake and distribution in the body

(i) Distribution in the blood

After absorption from the gastrointestinal tract, TCDD enters the lymph in the form of chylomicrons (Lakshmanan et al., 1986). Once in the blood, TCDD-containing chylomicrons are quickly (within 1 h) cleared from the blood. Cleared TCDD appeared mainly in the liver and the adipose tissue (74–81% of the administered dose). After clearance of chylomicrons, dioxin-like compounds remain mainly in serum lipoproteins (very low-, low- and high-density lipoproteins) and bound to serum proteins. In serum, the distribution of TCDD between lipoproteins and serum proteins is determined by their lipid content. However, higher-substituted dioxins and furans do not partition only in accordance with the lipid content of serum components: whereas the lipid content of serum lipoproteins is twofold higher than that of serum proteins, about 80% of TCDD resides are in serum lipoproteins and 20% in serum proteins. For OCDD, almost the opposite situation was observed, i.e. 40% in lipoproteins and 60% in serum proteins (Patterson et al., 1989). Furthermore, substantial partitioning of 1,2,3,6,7,8-HeCDD and 1,2,3,4,6,7,8-HpCDD between the serum and erythrocytes has been found, again indicating substantial binding of higher-chlorinated congeners to blood proteins.

(ii) Exchange between blood and organs

As in blood, the distribution of dioxins and furans between serum and organs is determined by lipid partitioning and protein binding. The concentrations of dioxins and furans in blood and adipose tissue correlate well (Päpke et al., 1989; Iida et al., 1999a). TCDD is distributed between plasma/blood and adipose tissue by lipid partitioning (Patterson et al., 1988; Gochfeld et al., 1989). However, in the case of HeCDD/HeCDF and OCDD/OCDF, the distribution between plasma and adipose tissue is determined by both lipid partitioning and plasma protein binding (Patterson et al., 1989; Schecter et al., 1991, 1998).

(iii) Hepatic sequestration in rodents

In the liver, protein binding plays an important role in the uptake of dioxin-like compounds from the blood, even for lower-chlorinated congeners. When rodents are exposed to increasing doses of TCDD, preferential accumulation occurs in the microsomal fraction of the liver, such that the concentration exceeds that in adipose tissue by many fold (Allen et al., 1975; Kociba et al., 1978a,b; Gasiewicz et al., 1983; Abraham et al., 1988; Leung, 1990a,b; Weber et al., 1993; Diliberto et al., 1996; Santastefano et al., 1996; Viluksela et al., 1996; Diliberto et al., 1999). The biochemical mechanism behind this phenomenon is as follows. After entering the liver cells, TCDD may dissolve in hepatic lipid, bind to an intracellular Ah receptor protein, or bind to cytochrome P450 (CYP) proteins, in particular CYP 1A2 (Poland et al., 1989a,b; Santastefano et al., 1996; Diliberto et al., 1997, 1998, 1999). As the amount of cellular CYP proteins is regulated by formation of the TCDD–Ah-receptor complex, exposure to increasing amount of TCDD triggers the cascade of events (protein induction) comprising increased entry of TCDD into the cell, increased formation of the TCDD–Ah-receptor complex, increased formation of CYP 1A2 mRNA and CYP 1A2 protein and increased binding of TCDD to the induced CYP 1A2 proteins (Whitlock et al., 1997).

Hepatic sequestration has also been observed with 2,3,7,8-TCDF and higher-chlorinated PCDDs and PCDFs (Yoshimura et al., 1984; Wacker et al., 1986; Couture et al., 1988; Abraham et al., 1989; Poiger et al., 1989a; DeVito et al., 1998; Diliberto et al., 1999) and PCBs (van Birgelen et al., 1994a,b). In the case of PCBs, hepatic sequestration depends on substitution at the ortho position, greater substitution resulting in decreasing sequestration. For example, 2,2΄,4,4΄,5,5΄-HxCB, 2,3,3΄,4,4΄-PeCB, 2,3΄,4,4΄,5-PeCB and 2,3,3΄,4,4΄,5-HxCB are preferentially deposited in adipose tissue and not in liver (van Birgelen et al., 1994a, 1995a; DeJongh et al., 1995; van Birgelen et al., 1996a; DeVito et al., 1998; Diliberto et al., 1999). In contrast, 3,3΄,4,4΄,5,5΄-HxCB and 3,3΄,4,4΄,5-PeCB reached relatively high concentrations in the liver and interfered with hepatic sequestration of TCDD (van Birgelen et al., 1994b).

The hepatic sequestration of dioxins, furans and PCBs markedly affects the relative amounts of these compounds in the body (body burden). For example, whereas the liver and adipose tissue contain 10% and 60% of the body burden of TCDD, respectively, in mice that have only constitutive hepatic CYP protein levels, the fractions may increase and decrease to 67% and 23%, respectively, in mice with induced hepatic CYP protein (Diliberto et al., 1995) and to 30% and 42% in rats (Diliberto et al., 1996).

Binary mixtures of dioxins, furans and PCBs show clear interactions with respect to hepatic sequestration. Co-administration of 2,2΄,4,4΄,5,5΄-HxCB with 2,3,3΄,4,4΄,5-HxCB doubled the hepatic disposition of the latter congener. A similar effect was found with co-administration of 2,2΄,4,4΄,5,5΄-HxCB and 3,3΄,4,4΄,5,5΄-HxCB (De Jongh et al., 1993a). The hepatic disposition of 1,2,3,7,8-PeCDD increased when administered with 2,2΄,4,4΄,5,5΄-HxCB, 1,2,3,6,7,8-HxCDD and 2,3,4,7,8-PeCDF (De Jongh et al., 1993b).

(iv) Hepatic sequestration in humans

Preferential sequestration of dioxins and furans in liver rather than adipose tissue has also been observed in persons exposed to background concentrations of these compounds (Figure 1). The observed hepatic sequestration is probably due to binding to constitutive rather than induced CYP 1A2 proteins, as, in humans, CYP 1A2 is primarily expressed constitutively and induced in the liver (Diliberto et al., 1999). Furthermore, although Ah receptor-dependent CYP induction has been observed in vitro in human liver cells exposed to TCDD (Schrenk et al., 1995; induction starting at 1 pmol/L; median effective concentration, 100 pmol/L), it occurred at concentrations that were several orders of magnitude higher than those observed in human blood, the mean TCDD concentration in human blood being 0.016–0.078 pmol/L (Päpke et al., 1989; Iida et al., 1991a,b; Schecter et al., 1991; Päpke et al., 1996; Schecter et al., 1998a). A physiologically based pharmacokinetics model showed that induction of Ah receptor-dependent CYP proteins is unlikely to occur in the liver of persons who have been exposed for long periods to background concentrations of TCDD (Zeilmaker et al., 1999).

FIGURE 1

Concentrations from Leung et al. (1190c) and Thomas et al. (1990). T4cdd, sum of tetracholodibenzodioxins; P5cdd, sum of pentachlorodibenzodioxins; H6cdd, sum of hexachlorodibenzodioxins; H7cdd, sum of heptachlorodibenzodioxins; Ocdd, octachlorodibenzodioxin; T4cdf, sum of tetrachlorodibenzofurans; P5cdf, sum of pentachlorodibenzofurans; H6cdf, sum of hexachlorodibenzofurans; H7cdf, sum of heptachlorodibenzofurans; Ocdf, octachlorodibenzofuran

Figure 1. Ration of concentrations of dioxins and furans in human liver and adipose tissue

(c) Elimination

Rodents excrete dioxins and furans almost exclusively via the bile, the urine being only a minor route of elimination (Gasiewicsz et al., 1983; Birnbaum, 1986; Poiger & Schlatter, 1986; Pohjanvirta et al., 1990; Diliberto et al., 1999). Whereas only the parent compound is found in the organs of rodents (Brewster & Birnbaum, 1987; Kedderis et al., 1991), mainly dioxin and furan metabolites occur in the bile (Birnbaum et al., 1980; Decad et al., 1981). The metabolism includes dechlorination, hydroxylation and conjugation (Koshakji et al., 1984; Wroblewsky & Olson, 1985; Pluess et al., 1987; Poiger et al., 1989a; van den Berg et al., 1994). Similar reactions have been found in human liver in vitro, with CYP 1A1 metabolism of 2,3,7,8-TCDF (Tai et al., 1993) and CYP 2B metabolism of 2,2΄,5,5΄-TCB (Ishida et al., 1991).

Excretion of unmetabolized dioxins and furans in faeces is an important route of elimination in humans, the contribution of faecal elimination to total elimination ranging from 14% (1,2,3,4,6,7,8-HpCDD) to 90% (OCDD) (Rohde et al., 1999). These findings suggest that some PCDDs and PCDFs are eliminated through metabolism in humans (van der Molen, 1998, 2000).

In rodents, the terminal half-life of TCDD is 8–24 days in mice (Gasiewicz et al., 1983; Birnbaum, 1985) and 16–28 days in rats (Rose et al., 1976; Koshakji et al., 1984; Abraham et al., 1988; Pohjanvirta et al., 1990; Weber et al., 1993). Humans eliminate dioxins and furans much more slowly than rodents. In one volunteer, the half-life of TCDD ranged from 5.8 to 9.7 years (Poiger & Schlatter, 1986; Schlatter, 1991). A half-life of 8.2 years was found in victims of the Seveso accident (Needham et al., 1994), and a half-life of 8.6 years was found in former chemical plant workers (Rohde et al., 1999).

Longitudinal, relatively extensive data showed a mean half-life for TCDD of 8.7 years in veterans of the Viet Nam war (Michalek et al., 1996) and 7.2 years in former workers in a herbicide plant (Flesch-Janys et al., 1996). In these analyses, first-order kinetics was used to estimate the half-lives from the time-dependent decrease in its concentration in blood. This approach is based on the assumptions that the body composition of individuals is constant during the observation period, that elimination is independent of body composition, and that individuals have a constant (background) rate of intake. Both groups of authors found that these assumptions were false. In order to correct for them, van der Molen (1998) and van der Molen et al. (2000) used a physiologically based pharmacokinetics model to calculate the half-life of TCDD from the data sets. This analysis resulted in in a half-life of 5 years in young adults, 11 years in elderly men and 8 years in 45-year-old men from the data of Michalek et al. (1996) and 4 years in young adults, 8.5 years in elderly men and 6.3 years in 45-year-old men from the data of Flesch-Janys et al. (1998). Thus, Michalek et al. (1996) found a value of 8.7 years compared with 8 years in the model, and Flesch-Janys et al. (1996) found a value of 7.2 years compared with 6.3 in the model. The mean half-life of TCDD in middle-aged men is thus 7.6 years. The reported half-lives of PCDDs, PCDFs and PCBs other than TCDD are shown in Table 2.

Table 2. Elimination half-lives for polychlorinated dioxins, furans, and coplanar polyclorinated biphenyls

Compounda

Half-time (range of means)

References

Polychlorinated dibenzodioxins and polychlorinated dibenzofurans

TCDD

4.0b–11

Poiger & Schlatter (1986), Schlatter (1991), Flesch-Janys et al. (1994), Needham et al. (1994), Flesch-Janys et al. (1996), Michalek et al. (1996), van der Molen (1998), van der Molen et al. (2000)

1,2,3,7,8-PeCDD

5.3b –16

van der Molen et al. (2000), Flesch-Janys et al. (1996), Rohde et al. (1999)

1,2,3,4,7,8-HxCDD

5.0b–14

van der Molen et al. (2000), Flesch-Janys et al. (1996), Rohde et al. (1999)

1,2,3,6,7,8-HxCDD

3.5–14

van der Molen et al. (2000), Flesch-Janys et al. (1996), Rohde et al. (1999), Gorski et al. (1984)

1,2,3,7,8,9-HxCDD

3.0b–7.3

van der Molen et al. (2000), Flesch-Janys et al. (1996), Rohde et al. (1999)

1,2,3,4,6,7,8-HpCDD

2.1b–4.4

van der Molen et al. (2000), Flesch-Janys et al. (1996), Rohde et al. (1999), Gorski et al. (1984)

OCDD

2.9b–8.3

van der Molen et al. (2000), Flesch-Janys et al. (1996), Rohde et al. (1999), Gorski et al. (1984)

2,3,7,8-TCDF

1.5b–3.2

van der Molen et al. (2000)

1,2,3,7,8-PeCDF

2.5b–5.3

van der Molen et al. (2000)

2,3,4,7,8-PeCDF

2.1c
5.0b-19.6

van der Molen et al. (2000), Flesch-Janys et al. (1996), Rohde et al. (1999), Ryan et al. (1993a)

1,2,3,4,7,8-HxCDF

2.6c
3.6b–8.1

van der Molen et al. (2000), Flesch-Janys et al. (1996), Rohde et al. (1999), Ryan et al. (1993a)

1,2,3,6,7,8-HxCDF

4.6b–9.5

van der Molen et al. (2000), Flesch-Janys et al. (1996), Rohde et al. (1999)

2,3,4,6,7,8-HxCDF

2.0b–11

van der Molen et al. (2000), Flesch-Janys et al. (1996), Rohde et al. (1999)

1,2,3,4,6,7,8-HpCDF

2.3c
1.8b–3.8

van der Molen et al. (2000), Rohde et al. (1999), Ryan et al. (1993a)

1,2,3,4,7,8,9-HpCDF

3.2b–6.9

van der Molen et al. (2000), Flesch-Janys et al. (1996)

OCDF

1.0b-2.1

van der Molen et al. (2000)

‘Non-ortho’ polychlorinated biphenyls

3,3΄,4,4΄,5,5΄-HxCB

10d

Ryan et al. (1993a)

‘Mono-ortho’ polychlorinated biphenyls

2,3,3΄,4,4΄-PeCB

0.6–3.9

Brown et al. (1989), Chen et al. (1982)

2,3΄,4,4΄,5-CB

0.3–5.8

Ryan et al. (1993a), Brown et al. (1989), Chen et al. (1982), Buhler et al. (1988)

2,3,3΄,4,4΄,5-CB

4.2

Ryan et al. (1993a)

a For abbreviations, see Table 1.

b Young adults

c Yu-Cheng patients, probably induced metabolism

d Based on only one case

Coplanar PCBs may induce their own metabolism. In rodents exposed to relatively high doses of 2,3,4,7,8-TCDF and TCDD, a twofold induction of their metabolism was observed (Brewster & Birnbaum, 1987; Leung et al., 1990b; McKinley et al., 1993). Similarly, clear biphasic elimination of 2,3,4,7,8-PeCDF and 1,2,3,4,7,8-HxCDF was observed in Yu-cheng and Yusho patients, indicating that they had been exposed to concentrations well above background for induction of metabolism (Ryan et al., 1993a). The metabolism of TCDD was found to be substantially induced in two patients with TCDD poisoning, the half-lives being 200 and 230 days (Geusau et al., 1999).

(d) Transport across the placenta

TCDD readily crossed the placenta of pregnant Long-Evans rats given a single oral dose of TCDD at 1.2 ΅g/kg bw in corn oil on day 8 of gestation. The concentrations of TCDD found in the fetal compartment (fetuses plus their placentae) were 39 pg/g (0.01% of the administered dose) on day 9 of gestation, when the maternal blood concentration was 15 pg/g; 26 pg/g (0.11% of the administered dose) on day 16, with a maternal blood concentration of 18 pg/g; and 21 pg/g (0.7% of the administered dose on day 21, with a maternal blood concentration of 8 pg/g. In individual embryos, the TCDD concentrations were 40, 18 and 22 pg/g on days 9, 16 and 21. The embryo/fetal compatment may therefore be considered a nonsequestering maternal compartment (Hurst et al., 1998).

Pregnant Long Evans rats received a single oral dose of TCDD at 0.05, 0.2, 0.8, or 1 ΅g/kg bw in corn oil on day 15 of gestation. On day 16, these doses resulted in concentrations of 6.8, 15, 50 and 61 pg/g in the fetal compartment and 5.3, 13, 39 and 56 pg/g in single, whole fetuses, with associated maternal body burdens of 31, 97, 520 and 580 pg/g. On day 21 of gestation, the concentration of TCDD were 4.3, 14, 32 and 37 pg/g in the fetal compatment, 4.3, 15, 32 and 36 pg/g in single, whole fetuses and 27, 76, 330 and 430 pg/g in the maternal body. On day 16 of gestation, there was a good correlation between the fetal and maternal body burden and the fetal body burden and maternal blood concentration, suggesting that, at a critical time, maternal blood concentrations provide an estimate of the concentrations of dioxin in the developing fetus. On day 16, 60% of the administered dose was recovered in the dams (Hurst et al., 2000a).

Long-Evans rats were given TCDD repeatedly before (5 days/week for 13 weeks) and during gestation at a dose of 1, 10, or 30 ng/kg bw in corn oil. On day 16 of gestation, the concentration of TCDD in single fetuses was 1.4, 7.8 and 16 pg/g, and the associated maternal body burdens were 19, 120 and 300 ng/kg bw, respectively (Hurst et al., 2000b).

2.1.2 Physiologically based pharmacokinetics modelling

As described above, the toxicokinetics of dioxin-like compounds involves the complex interaction of absorption, transport via the blood, distribution in the lipid and protein fractions of the blood and organs, Ah receptor-dependent induction of hepatic CYP proteins and elimination from the body by metabolism and/or transfer to faecal lipid. These concomitant processes can be quantified by physiologically based pharmacokinetics modelling, in which the toxicokinetics of chemicals is described mathematically within a physiological context, i.e. organs connected by the bloodstream and organ-specific responses after exposure to a chemical (Figure 2).

FIGURE 2

From Zeilmaker & Van Eijkeren (1999)

Q, blood flow; V, volume; P, partition coefficient; C, concentration; CYP, cytochrome P450 enzyme; D0, administered amount; Fabs, fraction absorbed over the gut wall; Vmax, maximum possible metabolic rate; KM, Michaelis-Menten constant; tau, delay; kappa, intra-compartmental diffusion; b, blood; f, fat, i.e. the body's adipose tissue; s, slowly perfused compartment (e.g. resting muscle, skin.); r, richly perfused compartment (e.g. lungs, kidneys, spleen); h, hepatic compartment, i.e. the liver. Arrows indicate the direction of blood flow, into (arterial blood) and from (venous blood) the organs

Figure 2. Physiologically based pharmacokinetics model for dioxins and furans

The first physiologically based pharmacokinetics models of dioxin-like compounds (2,3,7,8-TCDF, King et al., 1983; TCDD, Leung et al., 1988) allowed for only linear kinetics, i.e. they described the accumulation of dioxin-like compound in the body as a process which depends linearly on the administered dose. However, these models could not describe the hepatic sequestration of dioxin-like compounds, in particular the binding of TCDD to induced hepatic CYP 1A2. This deficiency was overcome by introducing Ah receptor-dependent CYP induction and subsequent binding of TCDD to induced CYP 1A2 in the model. The latter model was found to describe well the non-linear kinetics of TCDD, as observed in rodents exposed to TCDD at doses that induce Ah receptor-dependent CYP proteins in the liver (Leung et al., 1990a,b; Andersen et al, 1993; Kohn et al., 1993, 1994, 1996; Andersen et al, 1997a,b; Wang et al., 1997a,b; Zeilmaker & van Eijkeren, 1997).

Physiologically based pharmacokinetics modelling has also been used to describe the kinetics of dioxin-like compounds in humans. Assuming that the body burden of dioxin-like compounds is composed mainly of the amounts in the liver and adipose tissue, Carrier et al. (1995a,b) modelled the non-linear kinetics of dibenzo-para-dioxins and dibenzofurans resulting from their preferential accumulation in human liver. The model was used to describe the kinetics of 2,3,4,7,8-PeCDF in Yu-Cheng patients (see section 2.3.2). van der Molen (1998) developed a generic physiologically based pharmacokinetics model to describe the accumulation of dibenzo-para-dioxins and dibenzofurans in the blood and maternal milk of persons who had been exposed to background concentrations of these compounds. Kreuzer et al. (1997) and Pollitt (1999) used this type of modelling to evaluate the effect of exposure to these compounds in mother’s milk on the long-term body burden. In both cases, the exposure was found to have only a limited effect on the long-term body burden. Zeilmaker and van Eijkeren (1997; 1998) and Zeilmaker et al. (1999) used a slightly different approach: a physiologically based pharmacokinetics model of TCDD in rodents was scaled to humans. Figures 3, 4 and 5 show typical simulations made with this model of the accumulation of TCDD in the human body as expected after a variety of exposure scenarios.

FIGURE 3

First two panels: Life-long intake function of TCDD in women, absolute amount (pg) and relative amount (pg/kg bw oer day), 1 pg/kg bw per day

Third panel: Accompanying time-course of blood concentration (pg/g lipid, blood concentration of TCDD being equal to the concentration in the lipid fraction of adipose tissue). The effect of a single bolus oral dose of 100 (middle line) or 1000 (upper line) pg/kg bw administered at the age of 30 years on the blood concentration of TCDD in women exposed for life to 1 pg/kg bw per day. Extra Y-axis: TCDD concentration in blood lipid.

Fourth panel: Accompanying time-course of TCDD half-life (years) in body

Model as in Zeilmaker & van Eijkeren (1998) (fraction absorbed: 1)

Figure 3. Physiologically based pharmacokinetics model of accumulation of TCDD

FIGURE 4

First panel: Effect of transplacental exposure on the body burden after life-long exposure to TCDD (see first two panels, Figure 3). Y-axis: Body burden of TCDD expressed in pg/kg bw. Upper curve: transplacental transport, lower curve: no placental transport.

Second panel: Effect of a single dose of 100 or 1000 pg/kg bw on body burden of TCDD in women exposed for life to TCDD (see first two panels, Figure 3). Extra bolus oral dose administered at age of 30 years. Y-axis: Body burden of TCDD expressed in pg.kg bw.

Third panel: As in second panel, effect of both doses on ratio of area under the curve (AUC) of concentration-time

Fourth panel: Effect of lactation on body burden of mother. Lactation started at age 25 years and lasted 6 months. Milk production: 600 ml/day; fat, 4%. The lower line shows the body burden when the partition coefficients of adipose tissue and of milk fat are the same, i.e. 800, the line in the middle when the partition coefficient of milk fat is about one quarter of the adipose tissue partition coefficient, i.e. 160, resulting in a milk fat concentration of TCDD of about 2.4 pg/g milk fat, and the upper line shows no lactation.

Fifth panel: Combined effects of transplacental exposure and lactation on body burden of infant. Lactation as in fourth panel.

Model as in Zeilmaker & van Eijkeren (1998) (fraction absorbed: 1)

Figure 4. Physiologically based pharmacokinetics model of accumulation of TCDD

FIGURE 5

First panel: Dose-response relationship of the daily dose of TCDD up to age 30 years and the resulting blood lipid concentration. Lower curve line: model as in Zeilmaker & van Eijkeren (1998), i.e. a model of inducible metabolism. Upper straight line: model without induction metabolism.

Bottom panel. Accompanying metabolic induction factor with respect to basal metabolism. Note the absence of induction of metabolism at background exposure.

Figure 5. Physiologically based pharmacokinetics model of accumulation of TCDD

2.1.3 Biochemical effects

Cells exposed to chemicals may respond by increasing the activity of metabolizing enzymes, in particular phase I and phase II enzymes (enzyme induction). Although this mechanism can lead to the removal of chemicals with deletorious effects, enzyme induction also has clear disadvantages. As the induced enzymes often have broad substrate specificity, increased activity may increase the metabolism of chemicals other than the inducing compound. In particular, exposure to persistent chemicals like PCDDs, PCDFs and PCBs can lead to sustained, unwanted changes in chemical metabolism. Examples of the latter effect are the increased metabolism of thyroid hormones after induction of UDP-glucuronosyl transferase (UGT1) activity (see section on Thyroid hormones) and increased estrogen metabolism in the liver (see section on TGF-alpha/EGF pathway).

The following working model prevails for enzyme induction by TCDD (Whitlock et al., 1997). After TCDD has diffused into the cell, it binds to the intracellular Ah receptor protein, which is maintained in its ligand inactivated state by complexation with heat shock protein (hsp)-90. After binding of TCDD, the TCDD–Ah receptor complex dissociates from the hsp-90 protein. This complex than translocates to the nucleus, where it combines with the Ah receptor nuclear translocator (Arnt) to a transcription factor, which may bind to a specific DNA enhancer site, the so-called xenobiotic responsive element. Concurrently, transcription factors bind to gene-specific promotor sites, thereby increasing gene transcription. In this way, TCDD may modulate the transcription of CYP 1A1 (Kedderis et al., 1991; Tritscher et al., 1992; DeVito et al., 1996), CYP 1A2 (DeVito et al., 1996), CYP 1B1, NAD(P)H: quinone oxidoreductase, gluthathione A-transferase Ya subunit and UGT (Whitlock et al., 1997). Increased concentrations of protein may manifest as increased enzyme activities, the activity of ethoxyresorufin O-deethylase (EROD) relating mainly to CYP 1A1 and that of acetanilide-4-hydroxylase and methoxyresorufin O-demethylase mainly to CYP 1A2 (De Jongh et al., 1995; DeVito et al., 1996).

(a) Induction of CYP 1A1 and CYP 1A2

In vivo

TCDD efficiently induced CYP 1A1 and CYP 1A in rats, in which constitutive and inducible expression of CYP 1A2 is observed only in the liver. After administration of a single dose of TCDD ranging from 1 to 3000 ng/kg bw per day, < 50-fold induction of hepatic EROD activity was observed. Statistically significant induction was observed even at 3 ng/kg bw per day (Abraham et al., 1988). In concordance with this result, single doses of 0.1–1 ng of TCDD led to significant induction of CYP 1A1 mRNA in rat liver. A good correlation was found between hepatic CYP mRNA and EROD activity (r2 = 0.93) (van den Heuvel et al., 1994).

Similarly, CYP 1A1 and CYP 1A2 were induced by 23- and 5-fold, respectively, in rats given repeated doses of 3.5–125 ng of TCDD (Tritscher et al., 1992). As shown in Figure 6, whereas the hepatic TCDD concentration increased linearly as a function of the administered dose, the induced protein concentrations increased non-linearly as a function of the hepatic concentration of TCDD, until a maximum was reached.

FIGURE 1

Data from Tritscher et al. (1992); model described by Zeilmaker and van Eijkeren (1997)

Ordinate: hepatic TCDD concentration (nmol/kg); abscissa: hepatic TCDD concentration (nmol/kg)

Figure 6. Physiologically based pharmacokinetics model simulation of the concentration of cytochrome P450 (CYP) 1A1 and CYP 1A2 in the liver of rats exposed for 30 weeks to TCDD at a dose of 50, 150, 500, or 1750 ng/kg bw twice a week

TCDD-induced EROD activity is not limited to the liver. When B6C3F1 mice were given TCDD orally at a dose of 1.5–150 ng/kg bw per day on 5 days per week for 13 weeks, the activities of both enzymes were induced in the lungs and the skin at the lowest dose, being 30 times higher than the basal hepatic EROD activity in the lungs and 140 times higher in the skin. The dose–response characteristics of EROD induction were similar in the two organs. The lowest dose also significantly increased the phosphorylation of Cdc2 cyclin-dependent kinase, a protein associated with the G2 to M phase transition of cells, in the liver but not in skin (DeVito et al., 1994).

Whereas induction of acetanilide 4-hydroxylase is limited to the liver (DeVito et al., 1994), a single dose of TCDD at 0.1, 1, or 10 ΅g/kg bw to mice induced EROD activity in a dose-dependent fashion in liver, lungs and skin, the EROD activity in the lungs and skin being 6% and 0.6% of the corresponding hepatic activity (Diliberto et al., 1995). Similar observations were made in rats, in which dose-dependent induction of EROD activity and the amount of CYP 1A1 was observed in the liver, lungs and kidneys of rats given a single dose of 0.1, 1, or 10 ΅g/kg bw. As in mice, the induced EROD activity in the lungs and kidneys represented only a small fraction of the corresponding hepatic activity (lungs, 0.8%; kidneys, 2.5%). In all three organs, a strong correlation was found between induced EROD activity and the amount of CYP 1A1 protein. Similar observations were made for hepatic methoxyresorufin O-demethylase activity, with a dose-dependent increase in activity in the liver, which correlated well with induced CYP 1A2 protein levels. As expected, hardly any CYP 1A2 protein was observed in the lungs or kidneys (Santastefano et al., 1996).

CYP 1A1 and CYP 1A2 can also be induced by compounds other than TCDD. Administration of 2,3,4,7,8-PeCDF at single a dose of 300 ΅g/kg bw to C57BL/6N and 129/Sv mice caused marked induction of EROD and acetanilide 4-hydroxylase activity in the liver and of EROD activity in the lungs. Similar effects were not found after administration of 2,2΄,4,4΄,5,5΄-HxCB at a dose of 36 mg/kg bw (Diliberto et al., 1999). Hepatic EROD and acetanilide 4-hydroxylase activities were induced in B6C3F1 mice treated orally on 5 days per week for 4 weeks with TCDD at 0.15 ΅g/kg bw per day, with 2,3,7,8-TCDF at 1.5 ΅g/kg bw per day, with OCDF at 150 ΅g/kg bw per day, with 3,3΄,4,4΄-TCB at 15 mg/kg bw per day, with 2,3,4΄,4,4΄,5-HxCB at 30 ΅g/kg bw per day or with 3,3΄,4,4΄,5-PeCB at 1.5 ΅g/kg bw per day. No induction was found with 1,2,3,7,8-PeCDF at 9 ΅g/kg bw per day or with the PCBs 2,3,3΄,4,4΄-PeCB at 3 mg/kg bw per day, 2,3,3΄,4,4΄,5΄-HxCB at 300 ΅g/kg bw per day, or 3,3΄,4,4΄,5,5΄-HxCB at 3 ΅g/kg bw per day. Although the absolute activity of EROD in the liver was 15-fold higher than that in the lungs, similar dose–response relationships were found in these organs. In skin, increased EROD activity was found only with TCDD, OCDF and the PCBs 3,3΄,4,4΄-TCB and 2,3,3΄,4,4΄,5-HxCB (DeVito et al., 1993). Administration of OCDD on 5 days/week for 13 weeks at a dose of 50 ΅g/kg led to significant induction of EROD in the livers of Fischer 344 rats (Couture et al., 1988).

Significant EROD induction was found in the liver and skin of B6C3F1 mice given 2,3,7,8-TCDF at 1500 ng/kg bw for 4 and 13 weeks (DeVito & Birnbaum, 1995).

After administration of a single oral dose of 2,2΄,4,4΄,5,5΄-HxCB at 91 mg/kg bw to C57BL/6J mice, no EROD induction was observed in the liver, but induction was observed with the PCBs 2,3,3΄,4,4΄,5-HxCB at 17 mg/kg bw and 3,3΄,4,4΄,5,5΄-HxCB at 2.1 mg/kg bw. The induction was potentiated by concomitant administration of 2,3,3΄,4,4΄,5-HxCB and 2,2΄,4,4΄,5,5΄-HxCB, whereas no such potentiation was found with 3,3΄,4,4΄,5,5΄-HxCB, 1,2,3,7,8-PeCDD, 1,2,3,6,7,8-HxCDD, or 2,3,4,7,8-PeCDF (De Jongh et al., 1993). A single dose of 1000 ΅mol of 2,2΄,4,4΄,5,5΄-HxCB doubled hepatic EROD and acetanilide 4-hydroxylase activity in C57BL/6J mice (De Jongh et al., 1995). Similarly, B6C3F1 mice given 2,2΄,4,4΄,5,5΄-HxCB at a single dose of 360 mg/kg bw had a 2.5-fold increase in hepatic EROD and MROD and 17-fold increase in that of pentoxyresorufin-O-depentylase (van Birgelen et al., 1996c).

Dietary administration of TCDD (14–1024 ng/kg bw per day), 3,3΄,4,4΄,5-PeCB (0.47–10.1 ΅g/kg bw per day), or 2,3,3΄,4,4΄,5-HxCB (81–729 ΅g/kg bw per day) to Sprague-Dawley rats clearly induced EROD activity. No such induction was observed with 2,2΄,4,4΄,5,5΄-HxCB at 0.7–6 mg/kg bw per day (van Birgelen, 1995b). The 95% CIs for the NOELs were 0.35–0.89 for TCDD-induced EROD activity and 0.55–3.8 ng/kg bw per day for that of acetanilide 4-hydroxylase (van Birgelen et al., 1995a).

Oral administration of 3,3΄,4,4΄,5-PeCB to B6C3F1 mice at a dose of 0.015–1.5 ΅g/kg bw per day induced EROD and acetanilide 4-hydroxylase activity in the liver and EROD activity in the skin and lungs (LOEL, 0.045–1.5 ΅g/kg bw per day). The activities of liver-specific enzymes were induced by 2,3,3΄,4,4΄,5-HxCB at 450 ΅g/kg bw per day, and the activity of EROD in skin and lung was induced by doses > 1500 ΅g/kg bw per day. 3,3΄,4,4΄,5,5΄-HxCB at doses up to 3 ΅g/kg bw per day did not induce enzyme activity. The LOEL for the induction of hepatic EROD activity was 3900 ΅g/kg bw per day with the PCB 2,3,3΄,4,4΄-PeCB and 300 ΅g/kg bw per day with 2,3΄,4,4΄,5-PeCB, whereas that with TCDD was 0.0015 ΅g/kg bw per day (DeVito et al., 2000).

In vitro

Induction of EROD by PCDDs and PCDFs in primary hepatocytes has been found to be a sensitive end-point. Half-maximum EROD activity was achieved by incubating the cells with 16 pmol/l of TCDD, 90 pmol/l of 1,2,3,7,8-PeCDD, 184 pmol/l of 1,2,3,4,7,8-HeCDD, 329 pmol/l of 1,2,3,7,8,9-HeCDD, 441 pmol/l of 1,2,3,6,7,8-HeCDD, 702 pmol/l of 1,2,3,4,6,7,8-HpCDD, or 3859 pmol/l of OCDD (Schrenk et al., 1991). TCDD also induced EROD activity in primary human hepatocytes, although with very different dose–response characteristics. Half-maximum EROD activity was observed in cells exposed to ~ 100 pmol of TCDD (Schrenk et al., 1995).

(b) Induction of CYP 1B1

Like CYPs 1A1 and 1A2, CYP 1B1 is induced by TCDD. When C57BL/6J and DBA/2J mice were given single intraperitoneal doses of TCDD ranging from 0.001 to 50 ΅g/kg bw, dose-dependent accumulation of CYP 1B1 mRNA was observed in the liver. C57BL/6J mice showed a significant increase in mRNA at doses as low as 0.1 ΅g/kg bw, and maximum induction (200 times background) was found at 10 ΅g/kg bw. A much steeper dose–response curve was observed with CYP 1A1 mRNA than with CYP 1B1 mRNA (increase from 0.01 ΅g/kg; maximum induction at 1 ΅g/kg). The estimated effective doses at which half the maximum inducibility of CYP 1A1 and CYP 1B1 was observed (ED50) were 0.08 and 1.3 ΅g/kg bw. Similarly, in DBA/2J mice, significant induction of CYP 1A1 and 1B1 mRNA was found at doses of 1 and 10 ΅g/kg bw, respectively. Again, the dose–response curve of CYP 1A1 induction was much steeper than that for CYP 1B1, the ED50 values for CYP 1A1 and CYP 1B1 induction being 1.5 and 3.4 ΅g/kg bw, respectively. The differences in susceptibility of C57BL/6J and DBA/2J mice are due to several mutations in the Ahrd allele of the Ah receptor, resulting in lower ligand binding affinity: C57BL/6J mice carry the Ahrb-1 allele, conferring relative high susceptibility to TCDD, and DBA/2J mice carry the Ahrd allele, conferring relatively low susceptibility to TCDD (Abel et al., 1996).

(c) UGT1, PAI2 and TGF-alpha

In rat liver, the induction of UGT1, PAI2 and transforming growth factor (TGF)-alpha mRNA clearly deviated from that of CYP 1A1 mRNA. Although the dose required to increase UGT1 mRNA (1 ΅g/kg) was much higher than that required to induce CYP 1A1 mRNA (1 ng/kg), doses of TCDD up to 100 ΅g/kg did not induce PAI2 or TGF-alpha mRNA (Van den Heuvel et al., 1994).

(d) TGF-alpha and EGF pathway

Epidermal growth factor (EGF) is a plasma membrane receptor which, after binding a specific ligand, functions as a signal tranducer regulating cellular proliferation. This effect is mediated by internalization of the ligand–receptor complex and then phosphorylation of intracellular targets by tyrosine kinase. The effects of TCDD on hepatic EGF and their relationship to hepatic carcinogenesis are as follows. TCDD is a known inducer of liver tumours in female, but not male, rats (Kociba et al., 1978). It also induces proliferation of hepatocytes and preneoplastic foci in intact, but not ovariectomized, female rats (Lucier et al., 1991), indicating an important role of estrogens in hepatocarcinogenesis. Furthermore, treatment with TCDD results in a dose-dependent decrease in the number of EGF binding sites in the liver (Lucier et al., 1991; Kohn et al., 1993, short-term exposure to 3.5–125 ng/kg bw per day), indicating internalization of the receptor after TCDD binding (Kohn et al., 1993) and induction of TGF-alpha, a ligand of the EGF receptor.

The interactions of TCDD, EGF, TGF-alpha and inducible CYP enzymes in the liver are shown schematically in Figure 7. The hepatic TCDD–Ah receptor induces not only CYP 1A1 and 1A2 but also expression of the TGF-alpha gene and/or post-transcriptional or post-translational TGF-alpha modifications. This expression may be enhanced by the estrogen receptor–estrogen complex. As a consequence, more TGF-alpha is secreted into the interstitial space in the liver, where it may combine with EGF on the liver cell membrane. The TGF-alpha–EGF receptor complex may then be internalized to exert its biochemical signalling function. The TCDD–Ah receptor complex inhibits synthesis of the estrogen receptor. Finally, estrogen metabolism (estradiol-2-hydroxylase activity) is catalysed by CYP 1A2.

FIGURE 1

From Kohn et al. (1993). E, estrogen; E2OH, hydroxylated estrogen; ER, estrogen receptor; Ah, aryl hydroxylase receptor; TGF, transforming growth factor; ER-E2, complex of estrogen receptor and estrogen; EGF, epidermal growth factor

Figure 7. Interaction of TCDD, CYP 1A1, CYP 1A2, TGF-alpha and EGF receptors

(e) Thyroid hormones

PCDDs, PCDFs and PCBs may affect plasma thyroid hormone levels in one of three ways:

(1) By induction of hepatic microsomal enzymes, resulting in accelerated metabolism and excretion in the bile (Bastomsky, 1977). The stimulation of thyroid metabolism may be compensated by increased amounts of thyroid-stimulating hormone (TSH) in the blood. When sustained, such compensation may result in chronic stimulation of the thyroid and, ultimatally, thyroid cancer (Kohn et al., 1996; Whitlock et al., 1997).

The main pathway is glucuronidation of thyroxine (3,5,3΄,5΄-tetra-iodothyronine, T4) by UGT1, which catalyses the formation of T4 glucuronides, which are excreted in the bile. At least two forms of glucuronosyltransferase contribute to the activity of T4 UGT: UGT 1A1 (known to be induced by 3-methylcholantrene) and UGT 1A2 (known to be induced by phenobarbital).

Decreased plasma T4 levels were found after exposure of rats to phenobarbital, 3-methylchlolanthrene or the PCB mixture Aroclor 1254. This decrease correlated well with increased T4 UGT activity (Barter & Klaassen, 1992). Furthermore, PCBs and phenobarbital caused increased biliary clearance of T4 (Bastomsky, 1974; McClain et al., 1989; Beetstra et al., 1991). Short-term dietary intake of TCDD (14–1000 ng/kg bw per day) by rats lowered the concentrations of total T4 and free T4 in plasma at doses > 47 ng/kg bw per day. No effect was found on total triiodothyronine (T3) (van Birgelen et al., 1995b). Furthermore, short-term dietary intake of TCDD (14–1000 ng/kg bw per day), 3,3΄,4,4΄,5-PeCB (0.47–10 ΅g/kg bw per day) or 2,3,3΄,4,4΄,5-HxCB (81–730 ΅g/kg bw per day) resulted in dose-dependen increases in the activity of hepatic UGT 1A1, T4 UGT and CYP 1A1 and concomitant decreases in free and total T4 in plasma. Total T4 in plasma and hepatic T4 UGT activity were negativily correlated, whereas UGT 1A1 and CYP 1A1 activities were positivily correlated. This suggests involvement of the Ah receptor in inducing hepatic T4 glucuroni-dation and, consequently, in modulating thyroid hormone metabolism (van Birgelen et al., 1994a,b, 1995a).

(2) By binding of hydroxy metabolites of PCBs to transthyretin, thereby affecting the binding of T4 (and retinol) to transthyretin, its major transport protein in blood (Brouwer & van den Berg, 1986)

(3) By a direct effect of PCBs on the functioning of the thyroid gland. For example, Aroclor 1254 inhibits proteolysis of thyroglobulin, the protein responsible for release of thyroid hormones from the gland (Collins & Capen, 1980). Furthermore, exposure to TCDD results in an increase in serum TSH concentration and, after long-term exposure, to an increased volume of thyroid follicular cells, increased thyroid weight and thyroid hyperplasia (Bastomsky, 1977; Andrae & Greim, 1992; Hill et al., 1989).

(f) Retinoids

In rats, short-term dietary intake of TCDD (14–1000 ng/kg bw per day) reduced the hepatic concentrations of retinol and retinylpalmitate at the lowest dose tested. The concentration of plasma retinol was increased concomitantly. The mechanism behind this effect consists of induction of CYP enzymes, which oxidize retinol and reduce the activity of acyl coenzyme A:retinyl acyltransferase and lecithin:retinol acyltransferase. Both enzymes participate in the esterification of retinol. In a short-term study in Sprague-Dawley rats treated in the diet, the NOEL of 2,3,3΄,4,4΄,5-HxCB for depletion of hepatic retinoids was 81 ΅g/kg bw per day (van Birgelen et al., 1994a). In a similar study, the LOEL of 3,3΄,4,4΄,5-PeCB for this effect was 0.47 ΅g/kg bw per day (van Birgelen et al., 1994b).

(g) Hepatic porphyria

Disturbed biosynthesis of haem may lead to porphyria, a condition in which precursors of haem accumulate in blood and are hence excreted in the urine.

Short-term intake by rats of diets containing TCDD (14–1000 ng/kg bw per day), 3,3΄,4,4΄,5-PeCB (0.47–10 ΅g/kg bw per day), 2,3,3΄,4,4΄,5-HxCB (81–730 ΅g/kg bw per day) or 2,2΄,4,4΄,5,5΄-HxCB (0.7–6 mg/kg bw per day) maximally induced a twofold increase in hepatic porphyrin concentration (lowest effect doses: TCDD, 47 ng/kg bw per day; 3,3΄,4,4΄,5-PeCB, 3.2 ΅g/kg bw per day; 2,3,3΄,4,4΄,5-HxCB, 360 mg/kg bw per day). Concomitant administration of TCDD and 3,3΄,4,4΄,5-PeCB or 2,3,3΄,4,4΄,5-HxCB resulted in an additional twofold increase in porphyrin concentration. When these compounds were administered with a non-inducing dose of TCDD (33 ng/kg bw per day), however, an 800-fold increase in accumulation of porphyrins was seen, in particular uroporphyrin III and heptacarboxylic porphyrin, in the liver, indicating a strong synergistic action. Furthermore, induced acetanilide 4-hydroxylase activity correlated well with induction of accumulation of porphyrins in the liver. The mechanism behind these effects consisted of CYP 1A2-mediated oxidation of uroporphyrinogen III to uroporphyrin III and induction of the activity of delta-aminolaevulinic acid synthetase, the rate-limiting enzyme in haem biosynthesis, by 2,2΄,4,4΄,5,5΄-HxCB (van Birgelen et al., 1996a).

Short-term intake of TCDD (0.15–450 ng/kg bw per day), 1,2,3,7,8-PeCDD (90–9000 ng/kg bw per day), 2,3,7,8-tetrabromodibenzo-para-dioxin (30–3000 ng/kg bw per day), 2,3,7,8-TCDF (15–1500 ng/kg bw per day), 1,2,3,7,8-PeCDF, 90–9000 ng/kg bw per day), 2,3,4,7,8-PeCDF, 9–900 ng/kg bw per day), OCDF (1.5–150 ΅g/kg bw per day), 3,3΄,4,4΄,5-PeCB (0.3 and 15 ΅g/kg bw per day), 2,3,3΄,4,4΄-PCB (390–39 000 ΅g/kg bw per day), 2,3΄,4,4΄,5-PeCB (3000–30 000 ΅g/kg bw per day) or 2,3,3΄,4,4΄,5-HxCB (45–4500 ΅g/kg bw per day) by mice resulted in hepatic porphyria in all cases (van Birgelen et al., 1996b). The relative potencies of these compounds to induce hepatic porphyrin accumulation were TCDD > PeCDD = tetrabromodibenzo-para-dioxin = 4-Pe-CDF a TCDF > 3,3΄,4,4΄,5-PeCB a 1-PeCDF > OCDF > 2,3,3΄,4,4΄,5-HxCB > 2,3΄,4,4΄,5-PeCB a 2,3,3΄,4,4΄-PeCB. Again, hepatic CYP 1A2 enzyme activity and total hepatic porphyrin accumulation correlated well.

The Ah receptor appears to play an important role in dioxin-like porphyria, as hepatic porphyria is associated with the inducibility of Ah hydroxylase after administration of TCDD to Ah-responsive C57Bl/6 and Ah-nonresponsive DBA mice (Jones & Sweeney, 1980). Furthermore, non-coplanar PCBs, such as 2,2΄,4,4΄,5,5΄-HxCB and 2,2΄,3΄,4,4΄,5,5΄-HpCB do not induce porphyria (Stonard & Greig, 1976; Van Birgelen et al., 1996b; Koss et al., 1993). Finally, hepatic porphyrin accumulation and Ah-receptor dependent benzo[a]pyrene hydroxylation (a well known Ah-receptor response) are increased by 2,2΄,3,3΄,4,4΄-HxCB and 2,2΄,3,4,4΄,5΄-HxCB (Stonard & Greig, 1976).

2.1.4 Toxic equivalency factors

Exposure to PCDDs, PCDFs and coplanar PCBs actually consists of exposure to a mixture of congeners with toxic effects (dermal toxicity, immunotoxicity, carcinogenicity, reproductive and developmental toxicity, disruption of endocrine functions) similar to those of TCDD, the most toxic congener of this class of compounds. This finding led to the development of the concept of toxic equivalency factors (TEFs). In this concept, the toxic potency of a congener, i.e. its toxicity as described in all studies in vivo and in vitro, is expressed relative to that of a reference compound, in this case TCDD (for which the TEF is arbitrarily set at 1).

Application of the TEF concept rests on the following assumptions (van den Berg et al, 1998), noting that the TEF concept does not apply to toxicity that is not mediated by the Ah receptor and does not take into account modulation of Ah receptor-dependent responses by non-Ah-receptor ligands:

Derivation of congener-specific TEFs is based on evaluation and interpretation of the results of all available studies of toxicity on the basis of expert judgement. The first consultations on TEF concluded that the concept can be applied to PCDDs and PCDFs (international TEFs; NATO/CCMS, 1988) and non-ortho and mono-ortho PCBs (Ahlborg et al., 1994). The TEF concept was re-evaluated in 1998 (van den Berg et al., 1998, 2000).

In attributing a TEF to a compound, the following criteria were used:

Furthermore, TEFs in mammals were derived mainly from studies conducted in vivo, in preference to data generated in vitro and/or quantitative structure–activity relationships. Studies of toxicity were placed in the following order of priority: long-term > short-term > acute. Similarly, Ah receptor-dependent toxic end-points were given priority over biochemical end-points such as enzyme induction. TEFs for mammals were considered to apply to humans too.

Currently, there is consensus on the TEFs for mammals listed in Table 1 (van den Berg et al., 1998, 2000; Scientific Committee on Food, 2000). When combined with data on residues in matrices such as tissue, soil and water, TEFs allow determination of the toxic equivalents concentration of the residue. For a particular residue containing a mixture of i PCDDs, j PCDFs and k PCBs, the toxic equivalence is calculated according to the following equation:

EQUATION 1

In using the TEF concept, it should be kept in mind that non-Ah receptor-mediated toxicity (decreased dopamine concentration, effects on retinoid and thyroid hormone concentrations and estrogen receptor binding), shown by some PCBs, is not covered. Similarly, the TEF concept does not apply to halogenated compounds other than PCDDs, PCDFs and PCBs which show Ah receptor-dependent toxicity (brominated and chloro or bromo analogues of PCDDs, PCDFs, naphthalenes, diphenyl ethers, diphenyl toluenes, phenoxyanisoles, biphenyl anisoles, xanthenes, xanthones, anthracenes, fluorenes, dihydroanthracenes, biphenyl methanes, phenylxylyl-ethanes, dibenzothiophenes, quaterphenyls, quaterphenyl ethers and biphenylenes). Furthermore, non-additivity was found with mixtures of PCDDs, PCDFs and PCBs. For example, antagonistic effects of non-coplanar PCBs have been described on Ah receptor-dependent effects like induction of EROD and fetal cleft palate in mice. However, synergistic interactions between PCBs and PCDDs/PCDFs have also been reported for effects such as those on CYP 1A1 and thyroid hormone concentrations (van den Berg et al., 1998). Whereas the currently agreed TEFs are based on toxicological evaluations of dose–response relationships between external exposure, i.e. the levels of intake of congeners, and toxicity in organs, future TEFs will be based on the dose–response relationship between internal exposure, i.e. the actual concentrations of congeners in tissues, and toxicity in organs (van den Berg et al., 1998).

2.2. Toxicological studies

2.2.1 Acute toxicity

The acute toxicity of TCDD and related dioxins and furans substituted in at least the 2, 3, 7 and 8 positions can vary widely between and among species (Table 3). For example, in guinea-pigs, an LD50 of 0.6 ΅g/kg bw was recorded after oral administration, as compared with an LD50 of > 5000 ΅g/kg bw in Syrian hamsters. Explanations for this variation include differences in the Ah receptor, such as size, transformation and binding to the dioxin response element, pharmacokinetics (metabolic capacity, tissue distribution) and body fat content (Geyer et al., 1990; van den Berg et al., 1994; Pohjanvirta et al., 1998). While data on acute toxicity were available for various commercial PCB mixtures (LD50 values usually > 100 mg/kg bw), few data were available on the acute toxicity of the individual coplanar PCB congeners in mammals. In Ah-responsive rodent species, it is thought that lethality correlates to Ah receptor binding affinity.

Table 3. Acute toxicity of TCDD

Species and strain

Sex

Route

LD50 (΅g/kg bw)

Reference

Mouse, B6
Mouse/D2A/2J
Mouse/B6D2F1

Male
Male
Male

Oral
Oral
Oral

180
2600
300

Chapman & Schiller (1985)

Rat, Sprague-Dawley

Male

Oral

43

Stahl et al. (1992)

Rat, Sprague-Dawley

Male
Female
Weanling male

Intraperitoneal

60
25
25

Beatty et al. (1978)

Rat, Hartley and Wistar

Male

Intraperitoneal

> 3000

Pohjanvirta & Tuomisto (1987); Pohjanvirta et al. (1998)

Rabbit, New Zealand white

Male
Female

Oral
Dermal

120
280

Schwetz et al. (1973)

Hamster, golden Syrian

Male
Female

Oral
Oral

1200
5100

Henck et al. (1981)

Guinea-pig, Hartley

Male

Oral

0.6–2.1

McConnell et al. (1978); Schwetz et al. (1973)

Mink

Male

Oral

4.2

Hochstein et al. (1988)

Chicken

Not reported

Oral

< 25

Greig et al. (1973)

Rhesus monkey

Female

Oral

50–70

McConnell et al. (1978)

One of the commoner symptoms associated with dioxin-induced death is generalized delayed wasting syndrome, characterized by inhibition of gluconeo-genesis, reduced feed intake and loss of body weight. Although some differences exist between species, other toxic responses observed after single doses of dioxins include haemorrhage in a number of organs, thymic atrophy, reduced bone-marrow cellularity and loss of body fat and lean muscle mass. For example, in groups of five or six golden Syrian hamsters of each sex given a single oral or intraperitoneal dose of TCDD at 0, 5, 25, 100, 250, 500, 750, 1000, 2000 or 3000 ΅g/kg bw, deaths occurred at doses > 500 ΅g/kg bw, oral administration generally being more toxic. At the highest dose, 33% of males treated intraperitoneally and 80% of those treated orally died. The consistent pathological findings included gradual loss of adipose and muscle tissue (wasting syndrome), thymic atrophy and gastrointestinal lesions. The estimated LD50 values were > 3000 ΅g/kg bw for intraperitoneal administration and 1200 ΅g/kg bw for oral dosing (Olson et al., 1980).

Groups of 11 male C57BL/6J mice that were responsive or sensitive to TCDD (Ahb/b) were given a single oral dose of TCDD at 0, 5, 100, 200, 300 or 400 ΅g/kg bw, and a congenic non-responsive strain (Ahd/d) was given a single oral dose of 0, 400, 800, 1600, 2400 or 3200 ΅g/kg bw. All mice were observed for 35 days before necropsy. Significant mortality occurred in groups of the Ahb/b strain given doses > 200 ΅g/kg bw, 100% of animals at the two higher doses dying by about 21 days after dosing; an LD50 value of 160 ΅g/kg bw was estimated by probit analysis. Conversely, few deaths were seen in the Ahd/d strain, only 33% of those at the high dose dying; an LD50 value of 3400 ΅g/kg bw was estimated. Decreased body-weight gain, increased liver weight and thymic and splenic atrophy were observed in both strains of mice (Birnbaum et al., 1990).

Groups of 30–60 female Sprague Dawley rats were given 1,2,3,4,6,7,8-HpCB at a total dose of 0, 2.5, 2.8, 3.1, 3.4, 3.8, 4.1, 5 or 10 mg/kg bw by gavage (four doses over 2 days) and then observed for deaths. Many deaths due to lethal wasting, haemorrhage and/or anaemia were observed at the four higher doses, 83–100% of the animals dying within 25 days after dosing. None of the animals at the lowest dose died, while 8.3%, 31% and 66% of those at the next higher doses, respectively, died by day 44 after dosing. The author noted that 30% (9/30) of animals at the lowest dose died from squamous-cell carcinoma of the lungs (Rozman, 1999).

2.2.2 Short-term studies of toxicity

The toxicity of TCDD and related coplanar chemicals in short-term studies is characterized, depending on the route and species, by similar biological and toxicological effects.

Mice

In studies of the porphyrinogenic potential of various chlorinated dioxins, furans and coplanar PCBs, groups of five female B6C3F1 mice were treated on 5 days per week for 13 weeks by gavage with concentrations related by TEFs to doses of TCDD of 0, 0.15, 0.45, 1.5, 4.5, 15, 45, 150 or 450 ng/kg bw per day: 1,2,3,7,8-PeCDD, TEF, 0.05; 2,3,7,8-tetrabromodibenzo-para-dioxin, TEF, 0.15; 2,3,7,8-TCDF, TEF, 0.3; 1,2,3,7,8-PeCDF, TEF, 0.05; 2,3,4,7,8-PeCDF, TEF, 0.5; OCDF, TEF, 0.003; 2,3,3΄,4,4΄-PeCB, TEF, 0.00001; 2,3΄,4,4΄,5-PeCB, TEF, 5 x 10–6; 2,3,3΄,4,4΄,5-HxCB, TEF, 0.0001; 3,3΄,4,4΄,5-PeCB, TEF, 0.03. Analysis of hepatic tissue for highly carboxylated porphyrins and CYP 1A1 and 1A2 induction indicated that the binding affinity of the congeners to the Ah receptor in vivo was related to CYP 1A2 induction, which was in turn correlated to hepatic porphyrin accumulation. The LOELs associated with significant increases in total hepatic porphyrin content were: 15 ng/kg bw per day for TCDF; 45 ng/kg bw per day for TCDD; 300 ng/kg bw per day for PeCDD, 2,3,4,7,8-PeCDF and 3,3΄,4,4΄,5-PeCB, 900 ng/kg bw per day for tetrabromodibenzo-para-dioxin and 1,2,3,7,8-PeCDF; 45 ΅g/kg bw per day for OCDF; 1500 ΅g/kg bw per day for 2,3,3΄,4,4΄,5-HxCB; 7500 ΅g/kg bw per day for 2,3΄,4,4΄,5-PeCB; and 13 000 ΅g/kg bw per day for 2,3,3΄,4,4΄-PeCB. Induction of hepatic porphyrins by the mono-ortho-substituted PCBs was greater than that estimated from the TEFs assigned to them, which was considered to be related in part to non-Ah receptor-dependent induction of CYP 2B1 and delta-aminolaevulinic acid synthetase (van Birgelen et al., 1996b).

Rats

Groups of 12 Sprague-Dawley rats of each sex given TCDD at 0, 0.001, 0.01, 0.1 or 1 ΅g/kg bw per day by gavage on 5 days/week for 13 weeks had decreased organ and body weights, haematological effects and deaths at the two higher doses. The deaths occurred only at the highest dose, four females dying during the 13 weeks of treatment and two males and two females dying between 14 and 49 days during the 13-week post-dosing observation period. Minor increases in relative liver weight (5–8%) observed in animals at 0.01 ΅g/kg bw per day were considered to be an adaptive response, as no corresponding histopathological changes were seen. The NOEL was thus 0.01 ΅g/kg bw per day (equivalent to 0.007 ΅g/kg bw per day when averaged over the 13 weeks), which resulted in TCDD concentrations in the liver of 2.6–3.7 ΅g/kg (Kociba et al., 1976).

Groups of six male and six female rats of the same strain were fed diets containing a variety of penta- and hexachlorinated dioxins and dibenzofurans, separately and in a mixture, for 13 weeks: TCDD at 0.2, 2 or 20 ΅g/kg bw per day; 2,3,4,7,8-PeCDF, 1,2,3,7,8-PeCDF and 1,2,3,6,7,8-HxCD at 2, 20 or 200 ΅g/kg bw per day; or a mixture of TCDD at 0.2 or 2 ΅g/kg bw per day, 1,2,3,7,8-PeCDD at 1 or 10 ΅g/kg bw per day, 2,3,4,7,8-PeCDF at 2 or 20 ΅g/kg bw per day and 1,2,3,6,7,8-HxCD at 1 or 10 ΅g/kg bw per day. On the basis of decreases in body weight, histological changes in the liver and thymus and deaths, the relative toxicity of each congener was seen to be related to its binding affinity to the Ah receptor, while the toxicity of the mixture concurred with an additivity concept (TCDD > 2,3,4,7,8-PeCDF > 1,2,3,6,7,8-HxCDF > 1,2,3,7,8-PeCDF > 1,2,3,4,8-PeCDF) (Poiger et al., 1989b).

Male and female Fischer 344 rats were given oral doses of TCDD designed to generate a liver concentration of 0.03, 30 or 150 ng/g. After an initial loading dose of 0.005, 2.5 or 12 ΅g/kg bw, three maintenance doses of 0.0009, 0.60 or 3.5 ΅g/kg bw were given every fourth day, and the animals were killed at various times up to day 14. The terminal body weights of males and females at the highest dose were significantly reduced by approximately 11%, while the relative liver weights were increased on average by 30% and 43% at the two higher doses, respectively. Induction of hepatic CYP 1A1 (all doses) and CYP 1A2 (two higher doses) was determined by northern blot hybridization, and dose-dependent induction of a human TCDD-responsive CYP gene was detected in the livers of rats of each sex. Induction of this gene has not been associated with various treatments designed to induce hepatocellular proliferation (Fox et al., 1993).

Two weeks after receiving a protocol designed to initiate preneoplastic lesions (N-nitrosomorpholine at 80 mg/l of drinking-water for 25 days), female Wistar rats, were given subcutaneous injections of various dioxins every 2 weeks for 13 weeks, at approximate daily doses as follows: TCDD, 0, 2, 20 or 200 ng/kg bw; 1,2,3,4,6,7,8-HpCDD, 0, 0.2, 2 or 20 ΅g/kg bw; or a defined mixture of 49 dioxin congeners, 0, 0.2, 2 or 20 ΅g/kg bw. At the end of treatment, the body weights of animals at the highest doses of all three treatments were decreased by 5–7%, and the relative liver weights of animals were increased at the two higher doses of TCDD and the highest doses of the other compounds. Hepatic EROD activity was significantly increased in all treated groups when compared with controls, with similar induction at the low, intermediate and high doses. Linear regression analysis of hepatic EROD activity and measured toxic equivalents (ng/g liver) revealed slight differences in the slope of the regression lines (m = 0.87, 0.76, 0.67 for TCDD, HpCDD and the dioxin mixture, respectively; all r2 = 0.92). When the tumour promoting ability of the three treatments was assessed (relative focal volume of ATPase-deficient preneoplastic liver tissue), similar results were obtained after modelling the toxic equivalents for liver with TCDD and the dioxin mixture; however, the response to the latter was about twofold lower at the highest dose. The authors concluded that TEFs based on enzyme induction in vitro provide only an approximation of the tumour promoting ability of dioxin congeners and gave an overestimate of the response to the HpCDD (Schrenk et al., 1994).

Groups of six male and six female Sprague-Dawley rats were treated by gavage with total doses of 1,2,3,6,7,8-HxCDF of 0, 18, 220, 1300, 4000 or 6000 ΅g/kg bw for females and 0, 31, 370, 2200, 6700 or 10 000 ΅g/kg bw for males over 13 weeks; the total dose was divided into four daily loading doses, each comprising about 13% of the total dose, and was followed by six maintenance doses given every 2 weeks, each equivalent to 7.8% of the total dose. Treatment induced a variety of biochemical and toxic effects similar to those seen with TCDD in a group of rats given a total dose of 42 ΅g/kg bw for females and 70 ΅g/kg bw for males by the same dosing regimen. On the basis of the measured end-points (deaths, hepatic EROD induction, decreased plasma T4 concentration, haematological indices), the TEF for this congener was estimated to be 0.007, in close agreement with the WHO-assigned TEF of 0.01 (Viluksela et al., 1997).

In an experiment of a similar design, groups of 20 rats of the same strain were given total toxic equivalents of 0, 0.14, 1.7, 10, 31 or 47 ΅g/kg bw for females and 0, 0.22, 2.6, 16, 47 or 70 ΅g/kg bw for males, with contributions of similar toxic equivalents from TCDD (TEF, 1), 1,2,3,7,8-PeCDD (TEF, 0.2), 1,2,3,4,7,8-HeCDD (TEF, 0.05) and 1,2,3,4,6,7,8-HeCDD (TEF, 0.007). The same regimen of four daily loading doses (each comprising about 10% of the total toxic equivalents) followed by six maintenance doses (each comprising about 10% of the total toxic equivalents) every 2 weeks for 13 weeks. Whereas there was a significant, dose-dependent increase in hepatic EROD activity even at the lowest toxic equivalents (0.14 and 0.22 ΅g/kg bw for females and males, respectively), most of the additional biochemical end-points (decreased hepatic phosphoenolpyruvate carboxykinase activity, serum glucose and serum total T4) were affected only by total toxic equivalents > 10 ΅g/kg bw. Overall, the effects seen with the toxic equivalent mixture, TCDD or 1,2,3,6,7,8-HxCDD alone were comparable (mortality rate, growth reduction, hepatic enzyme induction, haematological effects), providing support for the TEF and additivity concept for chlorinated dioxins (Viluksela et al., 1998a,b).

In an experiment designed to identify the lowest effective doses and body burdens of TCDD in rats, groups of eight female Sprague-Dawley rats were fed diets formulated to deliver TCDD at a dose of 0, 14, 26, 47, 320 or 1000 ng/kg bw per day for 13 weeks. The most sensitive effects, seen at the lowest dose, included hepatic CYP 1A1 and 1A2 induction and significant decreases in thymus weights and hepatic retinol concentration. At the higher doses, differences in liver, kidney and spleen weights and decreases in plasma T3 and free T4 concentrations were seen. The estimated NOEL (by sigmoidal curve fitting) for hepatic EROD induction was 0.35 ng/kg bw per day, which corresponds to a liver TCDD content of 0.037 ng/g (van Birgelen et al., 1995a,b).

Thyroid hormone status was assessed in rats in a short-term assay for tumour promotion. After initiation with N-nitrosodietheylamine at 70 days of age, female Sprague-Dawley rats were treated with TCDD by gavage every 2 weeks for 30 weeks at doses designed to deliver 0, 0.1, 0.35, 1, 3.5, 11, 36 or 125 ng/kg bw per day. The rats were necropsied 1 week after the last TCDD dose, and serum samples were analysed for T3, T4 and TSH. T4 concentrations were significantly reduced at doses of TCDD > 11 ng/kg bw per day in the initiated rats and > 36 ng/kg bw per day in the uninitiated animals (maximum reduction, 42%). While there was no effect on T3 concentration, that of serum TSH was increased by about 2.5-fold in uninitiated rats at the highest dose (3.3 ng/ml, with 1.3 ng/ml in controls). Histological changes in the thyroid included diffuse follicular hyperplasia; in rats at 3.5 ng/kg bw per day, the ratio of parenchymal to follicular area was significantly increased. Hepatic CYP 1A1 and UGT1 mRNA levels were increased at doses > 0.35 ng/kg bw per day and > 3.5 ng/kg bw per day, respectively (Sewall et al., 1995).

Guinea-pigs

Groups of 10 Hartley guinea-pigs of each sex were given diets containing TCDD at a concentration of 0, 2, 10, 76 or 430 ng/kg of diet for 13 weeks (equal to 0, 0.12, 0.61, 4.9 and 26 ng/kg bw per day for males and 0, 0.14, 0.68, 4.9 and 31 ng/kg bw per day for females). The effects induced were similar to those in rats, including deaths at the highest dose. The NOEL for changes in organ and body weights and clinical effects was 0.61 ng/kg bw per day, confirming the greater sensitivity of this species to TCDD (DeCaprio et al., 1989).

Rhesus monkeys

When eight female rhesus monkeys were given a diet containing TCDD at a concentration of 500 pg/g for 9 months, dermatological effects were seen by 3 months and changes in haemoglobin and erythrocyte volume fraction by 6 months, which persisted and increased in severity up to the end of treatment. One animal died after 7 months on the diet, and four further deaths occurred within 2 months after removal from the diet, after total TCDD intakes estimated to be 19 and 12 ΅g/kg bw per day (Allen et al., 1977). Similar effects were seen when three male rhesus monkeys were fed diets containing 2,3,7,8-TCDF at 5 or 50 ΅g/g for up to 6 months, except that the animals that survived to the end of treatment tended to recover quickly after being removed from the diets (McNulty et al., 1981).

2.2.3 Long-term studies of toxicity and carcinogenicity

Previous WHO expert groups have evaluated the carcinogenicity of PCDDs, PCDFs and PCBs (IARC 1987, 1997; van Leeuwen & Younes, 2000). Most of the long-term experiments designed to determine the toxicity of dioxin and coplanar chemicals were conducted with various rodent species or non-human primates. The carcinogenic effects assessed in long-term studies are summarized in Table 4.

Table 4. Results of bioassays for carcinogenicity

Species, strain

No.

Doses(΅g/kg bw per day)

Route

Duration

NOEL

LOEL

Tumours observed

Reference

Mouse, Swiss

45 M

0, 0.001, 0.1, 1

Gavage

1 year, 1 day/week

0.001

0.1

Hepatocellular carcinoma

Tóth et al. (1979)

Mouse, B6C3

43–50 M,
42–49 F

0, 0.36, 0.72

Gavage

52 weeks,
1 day/week

 

0.36

Hepatocellular adenoma or carcinoma

Della Porta et al. (1987)

Mouse, B6C3F1

50–75 M




50–75 F

0, 0.0014, 0.0071, 0.071



0, 0.006, 0.03, 0.3

Gavage

104 weeks,
2 days/week

0.0071




0.006

0.071




0.03

Increased incidence of hepatocellular adenoma or carcinoma
Increased incidence of thyroid follicular-cell adenoma and histiocytic lymphoma

National Toxicology Program (1982)

Rat, Osborne-Mendel

50–75 M




50–75 F

0, 0.0014, 0.0071, 0.071

Gavage

104 weeks,
2 days/week

0.0014




0.0071

0.0071




0.071

Increased incidence of thyroid follicular-cell adenoma or carcinoma
Increased incidence of neoplastic nodules in liver or hepatocellular carcinoma

National Toxicology Program (1982)

Rat, Sprague-Dawley

10 M

0, 0.00004, 0.00014, 0.0014, 0.014, 0.057, 0.29, 3.4, 34, 71

Diet

78 weeks,
7 days/week

0.00004

0.00014

Renal adenocarcinoma, skin and lung carcinoma, leukaemia

Van Miller et al. (1977)

Rat, Sprague-Dawley

50 M, 50 F

0, 0.001, 0.01, 0.1

Diet

2 years,
7 days/week

 

0.1

Hepatocellular carcinoma, squamous-cell carcinoma in lung and hard palate

Kociba et al. (1978b)

M, male; F, female

Mice

Groups of 45 male Swiss mice were given TCDD by gavage at a dose of 0, 0.007, 0.7 or 7 ΅g/kg bw per week, equal to 0.001, 0.1 and 1 ΅g/kg bw per day, for up to 1 year. Dose-dependent increases in the incidence of both ulcerative skin lesions and amyloidosis were observed at the two lower doses and a decreased lifespan in mice at the highest dose. An increased incidence of liver tumours (hepatomas and hepatocellular carcinomas) was observed at the intermediate dose, but a similar increase at the highest dose was not significant (Tóth et al., 1979).

In a study of the carcinogenicity of TCDD in two rodent species, groups of 50 Osborne-Mendel rats and 50 B6C3F1 mice of each sex were given TCDD by gavage twice a week for 104 weeks at a dose of 0, 0.01, 0.05 or 0.5 ΅g/kg bw per week for rats and male mice (equal to 0, 0.0014, 0.0071 and 0.071΅g/kg bw per day) and 0, 0.04, 0.2 or 2 ΅g/kg bw per week for female mice (equal to 0, 0.006, 0.03 and 0.3 ΅g/kg bw per day). Decreased body-weight gain was seen in male and female rats at the highest dose, and an increased incidence of hepatic lesions described as ‘toxic hepatitis’ was seen in both species at the highest dose. Increased incidences of follicular-cell adenomas or carcinomas of the thyroid were found in male rats at the two higher doses (16% and 22%, respectively), with a non-significant increase (13%) in female rats. Furthermore, 24% of female rats at the highest dose had neoplastic nodules in the liver and 6% had hepatocellular carcinoma. The incidence of hepatocellular carcinoma was also increased in male (34%) and female (13%) mice at the highest dose. Females at this dose had an increased incidence of follicular-cell adenomas of the thyroid (11%) and an increased incidence (with a dose-related trend) of histiocytic lymphomas in the haematopoietic system (National Toxicology Program, 1982).

Rats

Groups of 10 male Sprague-Dawley rats were maintained on diets containing TCDD at a concentration of 0, 0.001, 0.005, 0.05, 0.5, 1, 5, 50, 500 or 1000 ng/kg for up to 78 weeks, equal to weekly doses of 0, 0.0003, 0.001, 0.01, 0.1, 0.4, 2, 24, 240 and 500 ΅g/kg bw or daily doses of 0.00004, 0.00014, 0.0014, 0.014, 0.057, 0.29, 3.4, 34 and 71 ΅g/ kg bw. All animals at the five higher doses died, the time to 100% mortality ranging from week 3 for animals at 24 ΅g/kg bw per week to week 31 for animals at 0.4 ΅g/kg bw per week. Further deaths (40–50% of animals) occurred at 0.001, 0.01 and 0.1 ΅g/kg bw per week before the end of the study. Neoplasms were found at multiple sites in 57% of animals at doses > 0.0001 ΅g/kg bw per week. Among those reported were ear-duct carcinoma, renal adenocarcinoma, skin angiosarcoma, Leydig-cell adenoma, fibrosarcoma, squamous-cell carcinoma of the skin and lung, glioblastoma, astrocytoma, cholangiocarcinoma and lymphocytic leukaemia. No tumours were found at the lowest dose (Van Miller et al., 1977). The Committee noted that the small number of animals used and the high mortality rates limit interpretation of this study.

Groups of 50 Sprague-Dawley rats of each sex (86 rats of each sex as vehicle controls) were fed diets formulated with TCDD to provide a dose of 0, 0.001, 0.01 or 0.1 ΅g/kg bw per day for 2 years. Body weight and food consumption were measured throughout the study; haematological examinations and urinary analyses were performed on eight rats of each sex per group after 3, 12 and 23 months. Serum samples were collected twice during the study. The biochemical and histopathological examinations were extensive. Increased incidences of hepatocellular carcinoma, squamous-cell carcinomas of the lung, hard palate and tongue were observed at the highest dose. TCDD not only affected the incidence rates of cancer but had additional toxicological effects, particularly at the highest dose, which included increased mortality (females only), decreased body-weight gain, splenic and thymic atrophy, hepatic degeneration and necrosis. On the basis of increased urinary excretion of porphyrins and delta-aminolaevulinic acid and hyperplastic nodules in the liver in females at 0.01 ΅g/kg bw per day, the NOEL was 1 ng/kg bw per day, which, at termination, resulted in a concentration of TCDD of 540 ng/kg in fat and liver (Kociba et al., 1978b).

Rhesus monkeys

As part of a study of reproductive toxicity, groups of female rhesus monkeys were given diets containing TCDD at 5 or 25 pg/g diet for 3.5–4.0 years, providing doses of 0.15 and 0.67 ng/kg bw per day. Animals at the higher dose showed marginal signs of toxicity (Bowman et al., 1989; see section 2.2.5 for details).

2.2.4 Genotoxicity

Previous WHO expert groups have evaluated the genotoxicity of PCDDs, PCDFs and PCBs (IARC, 1987, 1997; van Leeuwen & Younes, 2000). Several short–term assays for genotoxicity with TCDD covering various end-points gave primarily negative results. Furthermore, TCDD did not bind covalently to mouse liver DNA.

In vitro

TCDD did not induce mutations in Salmonella typhimurium strain TA98, TA100, TA1535, TA1537 or TA1538 with or without the addition of an exogenous metabolic activation system (Geiger & Neal, 1981; Mortelmans et al., 1984). Assays for Tk+/– mutation in mouse lymphoma L5178Y cells gave variable results, the outcome depending on the mutant selection protocol used, methotrexate or thymidine selection leading to a positive response and ouabain or arabinose C selection leading to a negative response. Thioguanine selection resulted in a weakly positive response (Rogers et al., 1982; McGregor et al., 1991).

Sister chromatid exchange and micronuclei were found in human lymphocytes treated with TCDD, in the absence or presence of alpha-naphthoflavone (Nagayama et al., 1993, 1994).

In assays for cell transformation in C3H10T1/2 mouse and rat tracheal epithelial cells, TCDD increased the formation of foci in cells initiated with N-methyl-N-nitro-N-nitrosoguanidine (Abernethy et al., 1985; Tanaka et al., 1989). TCDD also transformed Ad12-SV40-immortalized cells but not primary human epidermal keratinocytes, as revealed by growth in soft agar, and increased foci formation, cell density and the carcinogenic response in nude mice. The transformed cells caused a 100% incidence of squamous-cell carcinomas when injected into nude mice and 0% in control mice (Yang et al., 1992).

TCDD did not induce unscheduled DNA synthesis in normal mammary epithelial cells (Eldridge et al., 1992).

OCDD did not induce mutations in S. typhimurium (Zeiger et al., 1988), and 1,2,3,6,7,8-HeCDD and 1,2,3,7,8,9-HeCDD did not transform C3H10T1/2 mouse cells (Abernethy & Boreiko, 1987).

In vivo

TCDD did not bind to mouse liver DNA (Turteltaub et al., 1990), but it induced DNA single-strand breaks in the liver and in peritoneal lavage cells in rats (Wahba et al., 1988, 1989; Alsharif et al., 1994).

When administered concomitantly with 12-O-tetradecanoylphorbol 13-acetate, TCDD increased the cell transforming capacity of peritoneal macrophages in mice in a dose-dependent manner (Massa et al., 1992).

TCDD enhanced alpha-naphthoflavone-induced sister chromatid exchange frequency in cultured rat lymphocytes (Lundgren et al., 1986). It did not induce sister chromatid exchange, micronuclei or chromosomal aberrations in mouse bone marrow (Meyne et al., 1985) or in lymphocytes of persons who had been exposed to high concentrations of TCDD (Reggiani, 1980; Tenchini et al., 1983; Zober et al., 1993).

TCDD increased the mutagenic and recombinogenic activity of N-ethyl-N-nitrosourea in the mouse spot test by twofold (Fahrig, 1993).

In lacI transgenic rats, TCDD increased neither the mutation frequency nor the mutation spectrum in the liver (Thornton et al., 2001). Similarly, TCDD did change the spontaneous spectrum of H-ras codon 61 point mutations in mouse liver, nor did it affect the mutation spectrum of H-ras mutations in hepatocellular adenomas and carcinomas after treatment of mice with vinyl carbamate (Watson et al., 1995).

2.2.5 Reproductive toxicity

(a) Multigeneration studies

Rats

In a three-generation study of reproductive toxicity, male and female Sprague-Dawley rats (16 males and 32 females in the control and high-dose groups; 10 males and 20 females at the low and intermediate doses) were maintained on diets containing TCDD designed to deliver a dose of 0, 0.001, 0.01 or 0.1 ΅g/kg bw per day. After 90 days on diet, F0 rats were bred to produce the F1a generation and then again, 33 days after weaning, to produce the F1b generation. The F1b and F2 litters were mated when the animals were about 130 days of age to produce the F2 and F3 generations, respectively. Fertility, litter sizes and neonatal survival were severely decreased for animals at the highest dose at the F0 matings and in ensuing generations at the intermediate dose. Although slight effects were seen on pup survival and renal morphology at the low dose, they did not occur consistently across all generations. The NOEL was 0.001 ΅g/kg bw per day (Murray et al., 1979). The steady-state body burdens of TCDD of the rats at the two lower doses were estimated to have been 0.029 ΅g/kg bw and 0.29 ΅g/kg bw, respectively (Peterson et al., 1993).

Groups of 30 pregnant Wistar rats were given a single dose of 3,3΄,4,4΄,5,5΄-HxCB at 0, 0.2, 0.6 or 1.8 mg/kg bw by gavage on day 1 of gestation, and their pups were followed through to breeding about 1 year later. In the F1 generation, fecundity and fertility were decreased at the highest dose, only 2/16 pairs resulting in a pregnancy, in comparison with 15/17 and 11/17 at 0.2 and 0.6 mg/kg bw, respectively. After a second breeding, in which 14 treated males and seven treated females at 1.8 mg/kg bw were mated to control animals, fertility was again reduced, with no pregnancies and possible effects on unspecified aspects of male reproductive behaviour (Smits-van Prooije et al., 1993).

(b) Single-generation studies

Mice

Groups of 10–12 female B6C3F1 mice were treated by gavage with various dioxin, furan and PCB congeners once every 3 weeks for five dosing periods to assess the survival of autotransplanted endometrial lesions. The doses used were 0, 1, 3 and 10 ΅g/kg bw for TCDD; 0, 3 and 30 mg/kg bw for 2,2΄,4,4΄,5,5΄-HxCB; 0, 100, 300 and 1000 ΅g/kg bw for 3,3΄,4,4΄,5-PeCB; 0, 10, 30 and 100 ΅g/kg bw for 2,3,4,7,8-PCDF; and 0, 2 and 20 mg/kg bw for 1,3,6,8-TCDD. Endometrial tissue was transplanted during the second dosing period. Hepatic EROD activity was induced in a dose-dependent manner at the two highest doses of TCDD and at all doses of 3,3΄,4,4΄,5-PeCB and PCDF. The diameter of the endometrial lesions was significantly increased only at 1 and 3 ΅g/kg bw of TCDD and 100 ΅g/kg bw of PCDF. Whereas the weights of the lesions appeared to decrease with increasing dose of TCDD, the weights were still higher than those of controls. No significant effects were seen on ovarian or uterine horn weights, but the thymus weights were decreased and the liver weights increased at the two higher doses of PCDF. Examination of ovarian tissue showed the absence of active corpora lutea and increased numbers of regressive corpora lutea in mice at 10 ΅g/kg bw TCDD and 100 and 1000 ΅g/kg bw 3,3΄,4,4΄,5-PeCB (Johnson et al., 1997). In a study of similar experimental design with TCDD only, a dose of 10 ΅g/kg bw significantly increased the diameter of the endometriotic site in female Sprague-Dawley rats, and doses of 3 and 10 ΅g/kg bw decreased the ovarian weight by 15% and 27%, respectively (Cummings et al., 1996).

Rats

Groups of 25 female Wistar rats were given an initial subcutaneous loading dose of [14C]TCDD at 0, 25, 60 or 300 ng/kg bw, followed by weekly maintenance doses of 0, 5, 12 or 60 ng/kg bw, beginning 2 weeks before the beginning of mating and continually through mating, gestation and lactation. The size of the maintenance doses was based on a reported elimination half-life of 3 weeks for adult rats. The dams were killed after weaning, and 20 male pups per group were either assessed at 70 or 170 days of age for sexual development (sex organ weights, sperm analysis) or bred to untreated females on postnatal day 170. The only reproductive or pregnancy outcome that was significantly affected was the pregnancy index (16% reduction from that of controls) and only at the highest dose (300 ng/kg bw loading plus 60 ng/kg bw maintenance). The numbers of sperm from the cauda epididymis and daily sperm production were decreased (the latter by up to 50%) and the epididymal sperm transit rates and per cent abnormal sperm increased at all doses on postnatal day 170. No effects were seen on reproductive performance (mating, pregnancy, fertility). At the highest dose, the serum testosterone concentration was decreased at adulthood, and permanent changes in the testicular tubuli were found, including pyknotic nuclei and the occurrence of cell debris in the lumen. Mounting and intromission latencies were significantly increased at the lowest and highest doses. The LOEL was 25 ng/kg bw loading plus 5 ng/kg bw maintenance, equivalent to 0.8 ng/kg bw per day, which, by the end of weaning corresponded to 0.24 ng/g liver in male pups (Faqi et al., 1998).

Groups of 5–10 immature female Sprague-Dawley rats were given TCDD at a single dose of 0, 0.3, 1, 3, 10, 30 or 60 ΅g/kg bw by gavage, and ovarian follicular maturation was induced 24 h later by a single intraperitoneal dose of 20 IU of equine chorionic gonadotropin. The rats were killed 24, 48, 56 and 72 h later (study termination), and body weight, ovarian weights, ovulation status and serum prolactin, luteinizing hormone and follicle-stimulating hormone were measured. In a separate experiment, rats were hypophysectomized before dosing with TCDD at 0, 3, 10 or 60 ΅g/kg bw, equine chorionic gonadotropin (same schedule) and a single dose of 10 ΅g bovine luteinizing hormone. Reduced body-weight gain and ovarian weights were seen in rats at doses > 10 ΅g/kg bw, and ovarian weights were reduced by about 50% in rats at the high dose at termination. Ovulation, as defined by the number of ova per rat, was also significantly reduced at doses > 10 ΅g/kg bw (maximum reduction, about 80% at 30 and 60 ΅g/kg bw). The serum concentrations of estradiol were significantly increased over control values at 10 and 60 ΅g/kg bw 48–56 h after administration of equine chorionic gonadotropin, the concentrations in rats at the high dose being more than four times higher than that of controls at termination. Treatment at 10 ΅g/kg bw also resulted in a 13–15-fold increase in the serum concentrations of luteinizing and follicle-stimulating hormone and suppression of the preovulatory surge induced in control animals 54 h after administration of equine chorionic gonadotropin. Similar decreases in body and ovarian weight and inhibition of ovulation were seen in the hypophysectomized animals treated with TCDD at 10 and 60 ΅g/kg bw (Li et al., 1995).

Non-human primates

Experiments in which rhesus monkeys where fed diets containing TCDD at 50 or 500 pg/g for 7 months resulted in reproductive failure: only two of eight monkeys at 50 pg/g and one of eight at 500 pg/g carried their infants to full term. It was estimated that the monkeys at 50 pg/g group had consumed 1.8 ΅g of TCDD during the 7 months before the initiation of breeding, equivalent to 1.5 ng/kg bw per day (Allen et al., 1979).

In relevant studies involving nonhuman primates, groups of eight female rhesus monkeys were fed diets containing TCDD at 0, 5 or 25 ng/kg for 3.5 or 4.0 years, respectively. After 7 months of treatment, the monkeys were bred to untreated males, produced the F1 generation and were bred again after 27 months on diet. While reproduction was not affected at 5 ng/kg, only 20% of the combined breedings of animals at 25 ng/kg diet produced viable offspring, while an average of 75–80% of breedings of controls and animals at 5 ng/kg succeeded (Bowman et al., 1989). The concentration of TCDD detected in mesenteric fat samples from weaned F1 infants of mothers at 5 ng/kg of diet at 5 months of age was 380 ± 140 pg/g tissue. The total amount of TCDD ingested by the mothers was estimated to have been 370 ng after 16.2 months on diet (Schantz et al., 1992). The body burdens of the maternal animals were estimated to be 25–42 ng/kg bw.

While there was no indication of maternal toxicity at the time, analysis of the monkeys 10 years after termination of the study revealed a dose-dependent increase in the frequency and severity of endometriosis, and three monkeys at the higher dose died due to severe peritoneal endometriosis. Endometrial lesions were detected in 33% of control monkeys and 71% and 86% of those at 5 and 25 ng/kg of diet, respectively (Rier et al., 1993). Moderate-to severe disease was seen only in TCDD-treated animals (43% and 71% at 5 and 25 ng/kg of diet, respectively). As there was a high concentration of coplanar PCBs in the blood of monkeys in which endometriosis was originally diagnosed, there may have been an unknown source of exposure to PCBs that caused the reported lesions (Rier et al., 2001a). The average dietary concentration of PCBs was reported to be 7.8 ΅g/kg (Schantz et al., 1992), which, on the basis of the values for TCDD intake, would represent an average intake of 0.2 ΅g/kg bw per day.

Groups of five or six female cynomolgus monkeys were given TCDD-containing gelatin capsules 5 days/week for 12 months after surgical auto-implantation of endometrial strips at multiple abdominal sites. The average doses of TCDD delivered were 0, 0.71, 3.6 and 18 ng/kg bw per day. Laparoscopic examinations were conducted on all monkeys at 1, 3, 6 and 12 months (study termination). At necrospy, significantly more endometrial strips were found to have survived in monkeys at the two higher doses than in controls (27% and 33% vs. 16%, respectively), while the size of the implants was increased only at the highest dose (maximum and minimum length and width measured). The surviving endometrial strips in animals at the lowest dose actually regressed in size. The serum concentrations of interleukin (IL)-6 were significantly decreased, while those of IL-6 soluble receptor were increased in monkeys at the highest dose at termination, with a 2.5-fold change for both cytokines (Yang et al., 2000).

(c) Developmental toxicity

Dioxins, and specifically TCDD, induce a distinct series of developmental effects, including fetal mortality, structural malformations and postnatal functional alterations, in a variety of species at doses below those associated with maternal toxicity. These effects are thought to be due in part to interactions with the Ah receptor and its related transcriptional factor, ARNT, the expression of which appears to be involved in aspects of normal embryonic development (Kozak et al., 1997; Abbott et al., 1999).

In most species tested to date, TCDD can induce significant embryolethality (early or late resorptions, abortions, stillbirths), which is usually associated with indications of maternal toxicity. The timing of dosing and the age of the embryo or fetus have been shown to be major determinants of TCDD-induced prenatal mortality. Characteristic, sensitive indications of the teratogenicity of TCDD in responsive strains of mice include induction of cleft palate and hydronephrosis at doses not associated with maternal toxicity. Various strains of pregnant mice treated with TCDD at doses of 1–10 ΅g/kg bw per day during days 6–15 of gestation (organogenesis) produced litters with these effects in a dose-dependent fashion (Neubert et al., 1973; for review, see Environmental Protection Agency, 2000a, Part II, Chapter 5). While cleft palate in the absence of fetal or maternal toxicity usually occurs only in mice, common effects in rats and hamsters include renal malformations, oedema and gastrointestinal haemorrhage. The ability of most coplanar chemicals to induce frank developmental effects is related to their binding affinity to the Ah receptor.

Female C57Bl/6N mice were given a single oral dose of TCDD at 0, 6, 12, 15 or 18 ΅g/kg bw on day 10 of gestation or 0, 6, 9, 12 or 15 ΅g/kg bw on day 12 and then killed before parturition on day 18. For comparison, mice were treated with retinoic acid at 0, 10, 20, 30, 40, 50 or 60 mg/kg bw on day 10 of gestation or at 0, 40, 80, 100, 120, 160 or 200 mg/kg bw on day 12. Groups of mice were also treated with various combinations of TCDD (at 0, 3, 6, 12 or 15 ΅g/kg bw) and retinoic acid (0, 5, 10, 20, 30, 40 or 80 mg/kg bw) on day 10 or 12 of gestation. Decreased maternal body-weight gain was seen only in mice treated with TCDD on day 10 of gestation, whereas the relative liver weight was increased at all doses of TCDD. There was a trend towards a decrease in maternal body-weight gain in mice treated with retinoic acid on day 10 and increased relative liver weight after treatment on both days 10 and 12. The increase in relative liver weight seen in mice treated with both chemicals was similar to that seen with TCDD alone. No effects were seen on the mortality rate or body weight of fetuses of TCDD-treated mice, while the body weights of fetuses of mice at the two higher doses of retinoic acid given on day 12 of gestation were reduced by 6–10%. All treated groups had 6–20 litters. Hydronephrosis was seen in all groups treated with TCDD, with a dose-dependent increase in severity. The median effective dose (ED50) for this lesion was estimated to be 3.9 ΅g/kg bw for both times of treatment. The incidence of cleft palate was also increased by TCDD at the three higher doses at both times of treatment, with approximate ED50 values of 12 ΅g/kg bw for treatment on day 12 and 16 ΅g/kg bw for treatment on day 10. Major defects observed in the mice treated with retinoic acid consisted of skeletal abnormalities and cleft palate, with ED50 values for the latter of 42 mg/kg bw for treatment on day 10 and 190 mg/kg bw for treatment on day 12. Combined treatment with the two chemicals had no effect on either TCDD-induced hydronephrosis or retinoic acid-induced skeletal anomalies, but they had an interactive effect on cleft palate induction. For example, at doses of retinoic acid that induced no cleft palate (5 and 10 mg/kg bw), concomitant administration of TCDD at 12 ΅g/kg bw increased the incidence by approximately 2.4- and 4.2-fold, respectively (21% per litter as compared with 50% and 88%) (Birnbaum et al., 1989).

Groups of 7–10 pregnant C57Bl/6 mice were given a single oral dose of 3,3΄,4,4΄,5-PeCB at 0, 130, 260 or 520 ΅g/kg bw on day 10 of gestation and were killed on day 17. Significant increases in relative liver weights were seen in maternal animals at all doses, but there were no effects on litter size, litter weight or fetal mortality. Dose-dependent increases in the incidence of both hydronephrosis and cleft palate were seen, with ED50 values of 160 ΅g/kg bw and 360 ΅g/kg bw, respectively (Mayura et al., 1993).

Various structural end-points associated with the developing urogenital system of rodents have been shown to be extremely sensitive to perturbation by TCDD and coplanar chemicals. In males in particular, indices of delayed puberty (rats and hamsters), reductions in prostate growth and development (rats and mice), decreased epididymal and cauda epididymal weights (rats), decreased testicular and epididymal sperm numbers (rats, mice, hamsters) and decreased numbers of ejaculated sperm (rats and hamsters) have been observed (Roman & Peterson, 1998). In certain strains of rat (Holtzman and Long Evans), TCDD given as single doses as low as 0.05–0.064 ΅g/kg by gavage on day 15 of gestation has been associated with reduced ventral prostate weights and reduced epididymal and ejaculated sperm counts (Mably et al., 1992a,b,c; Gray et al., 1995, 1997a,b).

Pregnant Holtzman rats were given a single oral dose of TCDD at 0 or 1.0 ΅g/kg bw on day 15 of gestation, and the litters were collected at various times up to day 21 or the pups were collected after parturition up to postnatal day 5. Additional maternal rats were treated with a single oral dose of 0, 0.064, 0.16, 0.4 or 1 ΅g/kg bw on day 15 of gestation, and the postnatal physical and sexual development of the offspring was followed up to postnatal day 120. While no maternal toxicity was seen, fetal survival was decreased by 6–15% and neonatal body weight was decreased by 7–22% at the highest dose. When the food intake of male offspring was assessed after lactation, it was found to be decreased at the two higher doses, by 13–21% from that on postnatal day 28–56. This contributed to a decrease in body weight of 8–19% in the same groups up to postnatal day 70 (post-pubertal). After this time and until sexual maturity (postnatal day 120), no significant differences were seen for either parameter.

With regard to physical development, eye opening was accelerated by approximately 1 day in animals at the highest dose when compared with controls, while testis descent (androgen-dependent) was delayed in a dose-dependent manner at doses > 0.16 ΅g/kg bw (maximum delay, 2 days). A decreased relative anogenital distance were seen on day 4 in male pups of dams at doses > 0.16 ΅g/kg bw. Treatment at 1 ΅g/kg bw also caused a significant decrease (53%) in the post-parturition testosterone surge normally seen in male pups within 2–4 h after birth. Although there was a trend towards decreased plasma testosterone and dihydrotestosterone concentrations in male offspring from postnatal day 32–120, none of the changes was significant. The plasma concentration of luteinizing hormone in males at the highest dose was significantly reduced by about 95% on postnatal day 32. The weights of the seminal vesicle and ventral prostate were reduced in male offspring of dams at doses > 0.16 ΅g/kg bw throughout sexual development, but at the highest dose only the latter effect persisted at sexual maturity (postnatal day 120) (Mably et al, 1992a).

Male offspring from litters in the above study, standardized on postnatal day 1 to five pups of each sex per litter from dams treated with graded doses of TCDD on day 15 of gestation, were also assessed for sexual behaviour on postnatal days 56–63, 70–77 and 112–119. The latencies to mounting and to ejaculation were significantly increased in pups of dams at doses > 0.16 ΅g/kg bw at all times. While the number of intromissions required before ejaculation was not significantly affected, the number of mounts required before ejaculation was slightly increased in all groups during testing on postnatal days 112–119. The copulatory rate (mounts plus intromissions per minute) was significantly decreased in a dose-dependent manner from that of controls at all times (maximum reduction, about 53%) in male offspring of dams at the three higher doses. Feminine sexual behaviour (lordosis) was also significantly increased in castrated male offspring of dams at doses > 0.16 ΅g/kg bw, while a feminized pattern of luteinizing hormone surge after progesterone treatment was seen in castrated males after treatment of dams at the two higher doses. Hepatic EROD activity, when determined on postnatal day 120, was not significantly increased at any dose (Mably et al., 1992b).

Male pups isolated after lactation were assessed for various aspects of the development of the reproductive system on postnatal day 32, 39, 63 and 120 (juvenile, pubertal, postpubertal and sexually mature, respectively). Although slight decreases in paired testis weights were observed at all doses except 0.16 ΅g/kg bw on postnatal day 32, no significant difference was apparent by postnatal day 120. The weight of the right epididymis was decreased in a dose-dependent manner at a maternal dose as low as 0.064 ΅g/kg bw on postnatal days 49 and 120 and 0.16 ΅g/kg bw on postnatal days 32 and 63, with a 22–30% decrease at the highest dose. The weights of the cauda epididymis were lower than those of controls at all doses after puberty (postnatal days 63 and 120). Daily sperm production and cauda epididymal sperm numbers were also decreased at all doses on postnatal days 63 and 120, those of pups of dams at the highest dose being as little as 74% and 44% of control values on postnatal day 120. No significant effects were seen on cauda epididymal sperm motility or morphology. The plasma concentration of follicle-stimulating hormone was slightly reduced only on postnatal day 32 at all maternal doses except 0.16 ΅g/kg bw. When the male offspring were mated with control females on postnatal day 70, no effects were observed on fertility, gestation index, litter size or pup survival (Mably et al., 1992c).

Groups of 12 pregnant Long Evans rats were given TCDD at a single dose of 0, 0.05, 0.2 or 0.8 ΅g/kg bw by gavage on day 15 of gestation, and maternal and pup development and viability were followed until postnatal day 63. Some male pups were retained for assessment of reproductive organ weights and sperm counts at 15 months of age. Both pup survival and body-weight gain by the end of lactation were significantly decreased (15–17%) in the group at the highest dose, and eye opening was accelerated by 1 day at all doses in comparison with controls. On postnatal days 49 and 63, most of the developmental affects in male offspring occurred at the highest dose and included decreased numbers of epididymal sperm (postnatal day 49 and 63) and decreased weights of the cauda epididyma (postnatal day 63), ventral prostate and seminal vesicle (postnatal day 49). The onset of puberty was delayed by about 2 and 4 days at the two higher doses, respectively. In males at 15 months of age, significant reductions in the weight of the glans penis (8%) and the numbers of epididymal and cauda epididymal sperm (11–17%) were seen at the two higher doses. The number of ejaculated sperm was decreased (47%) only at the highest dose. However, when the results were combined with those of a previous study (Gray et al., 1995), animals at both 0.05 and 0.2 mg/kg bw also showed a significant decrease (25–26%) (Gray et al., 1997a).

These effects of low doses of dioxin and coplanar chemicals on the developing male reproductive system are usually not paralleled by decreased reproductive capacity or androgen status. Also, in general, the effects of TCDD and related coplanar chemicals on the urogenital tract during fetal and neonatal development of rodents, although dependent on the timing and age of the fetus, can be achieved by low doses in utero alone.

Not only male reproductive tract development but also various aspects of male sexual behaviour can be affected by perinatal exposure to low doses of TCDD and coplanar PCBs. In the experiments by Mably et al., assessment of the sexual behaviour of male Holtzman rats on postnatal days 60, 75 and 115 showed increased latency for intromission, mounting and ejaculation in offspring of dams given doses of TCDD as low as 0.064 and 0.16 ΅g/kg bw, respectively. Similarly, in Long Evans rats (Gray et al., 1995), partial demasculinization of male sexual behaviour, as determined by an increased total number of mounts, mounts with intromissions before ejaculation and latency before ejaculation, was observed. Increased numbers of mounts with intromissions before ejaculation were also observed in male Wistar rats after maternal exposure to 3,3΄,4,4΄,5-PeCB at a dose of 10 ΅g/kg bw (Faqi et al., 1998). While demasculinization of sexual behaviour was reproducible in three strains of rat, it was not seen in hamster offspring. Feminization of male sexual behaviour, as assessed by increased intensity and duration of lordosis, was also observed in male Holtzman rats at maternal doses of TCDD as low as 0.16 ΅g/kg bw (Mably et al., 1992c) but not in Long Evans rats. Alteration of defeminization of sexual behaviour by TCDD has been shown to require exposure during lactation or after parturition.

In rats and hamsters, single maternal doses as low as 0.2 and 2 ΅g/kg bw on days 15 and 11.5 of gestation, respectively, also increased the frequency of female pups with cleft phallus and vaginal threads, the latter condition being specific to coplanar chemicals (Gray et al., 1997b; Wolf et al., 1999). However, in ICR mice, a single maternal dose of TCDD at < 60 ΅g/kg on day 14 of gestation was not associated with genital malformations in female offspring (Theobald & Peterson, 1997).

Female offspring from the study of Gray et al. (1997a) were isolated after weaning and followed for sexual development until postnatal day 70. Additional females were mated to control males on postnatal day 100 for assessment of fertility. On postnatal day 70, the age of onset of vaginal opening was delayed in pups of dams at 0.8 ΅g/kg bw by approximately 2 days when compared with controls, and pups of dams at 0.2 and 0.8 ΅g/kg bw had decreased distance between urethral and vaginal openings (by 40% at the highest dose). After mating on postnatal day 100, no significant differences were seen in treated animals with respect to fertility, but the time to pregnancy of offspring of dams at the highest dose was increased by three estrous cycles when compared with controls (4.1 days and 14.4 days, respectively). Significantly increased numbers of female offspring at 0.2 and 0.8 ΅g/kg bw had urogenital anomalies (vaginal thread, cleft phallus). For example, 92% of females at the highest dose and 2.5% of controls had permanent vaginal threads. A cross-fostering experiment with the group at 1 ΅g/kg bw indicated that exposure in utero was necessary for the induction of vaginal and external genitalia anomalies. Cleft phallus and changes in the urethral opening position have also been observed in rats given other estrogen-like chemicals (diethylstilbestrol, RU 2858, estradiol) (Gray et al., 1997b).

An increase in the severity of urogenital malformations was seen in groups of eight pregnant Long Evans rats given TCDD at 1 ΅g/kg bw on day 15 of gestation as compared with day 8 of gestation. After parturition, the litters were standardized to four pups of each sex, and the female offspring were monitored for sexual development, estrous cycle and fertility. Although no maternal toxicity or effect on litter size were noted, the body weights of pups were significantly reduced (average, 17%) from postnatal day 3 throughout the study in the group treated on day 15 of gestation but only on postnatal day 3 for the group treated on day 8; pup survival was decreased by 11% in the group treated on day 15. Both treatments resulted in a variety of malformations in the external genitalia of female offspring, including complete or partial clefting of the phallus and the presence of a vaginal thread; the severity and/or frequency were usually greater in offspring treated on day 15 of gestation. For example, a vaginal thread was present in 79% of progeny treated on day 15 and 14% (nonsignificant) of those treated on day 8. When estrus was assessed throughout the reproductive lifespan, from postnatal day 125, more offspring treated on day 8 (47%) were in constant estrus at about 1 year of age than either controls (16%) or those treated on day 15 (16%). This effect was thought to be related to the decreased fertility rate of female progeny treated on day 8 when they were entered into a continuous breeding phase (five successive litters). Whereas the fertility rates of females treated on day 15 were similar to those of controls up to the fourth litter (but significantly reduced by the fifth litter), the fertility rate of females treated on day 8 was reduced by the second litter and throughout the duration of the breeding phase. Although overall fertility and fecundity in the first litter produced by females treated on day 15 were not affected, the high prevalence of vaginal threads caused greater difficulty in mating with control males, as illustrated by the increased number of mounts without intromission and increased ejaculation latency. Similar genital malformations (83% of offspring with vaginal threads) and suppressed body weight were seen in female offspring of Holtzman rats treated with TCDD at 1 ΅g/kg bw on day 15 of gestation, except that neonatal survival was decreased by 50% by postnatal day 16. Owing to the low pup survival, reproduction was not assessed in Holtzman rats, although the ovarian weights were reduced by about 47% in comparison with controls at necropsy (Gray & Ostby, 1995).

Pregnant Syrian hamsters, about 4 months old, were given a single dose of TCDD at 0 (n = 9) or 2 ΅g/kg bw (n = 10) on day 11.5 of gestation by gavage. The dose reflected the higher LD50 seen with this species than with rats (1200 ΅g/kg bw and 20 ΅g/kg bw, respectively). After parturition and lactation, the female offspring were monitored for various developmental landmarks until maturity (postnatal day 152–166). Although TCDD treatment had no effect on the body-weight gain or litter size of dams, pup body weights were reduced by up to 30% by the end of lactation, and the reduction persisted until postnatal day 140. Eye opening was accelerated by about 1 day in the pups treated with TCDD, while vaginal opening was delayed (100% of control pups by postnatal day 13 and 100% of TCDD-treated pups by postnatal day 19). After mating of the female offspring on postnatal day 152–166, a variety of parameters, including fertility, weaning index, number of implants, litter size and pup survival, were significantly reduced in the treated animals. TCDD treatment also increased the frequency of F1 female offspring with mild hypospadias or cleft phallus (0.78% in controls, 92% with TCDD) (Wolf et al., 1999).

Pregnant ICR mice (CD-derived) were treated by gavage on day 14 of gestation with TCDD at a single oral dose of 0, 15, 30 or 60 ΅g/kg bw (n = 17, 16, 19 and 14, respectively). The litters were standardized on postnatal day 1 to four of each sex, and development was assessed after interim sacrifices on postnatal days 23 (males only), 44, 65, 113–114, 128 (males only) and 142 (females only). Duration of gestation, prenatal mortality, litter size and dam weight were not affected by TCDD, and no effect was seen on postnatal pup mortality or pup weight gain throughout the experiment. Assessment of male reproductive organs during development indicated no effects on anogenital distance, onset of puberty, serum testosterone concentration, growth of the testis, epididymis, dorsal prostate or seminal vesicle, while the weights of the ventral prostate and coagulating gland (anterior prostate) were decreased at most doses from postnatal day 44 and continuing throughout the assessment (36% and 27% decrease at the highest dose, respectively, on postnatal day 114 or 128). While the epididymal sperm counts of mice at 30 and 60 ΅g/kg bw were reduced by about 30% on postnatal day 65, no effect was seen on postnatal day 114 or 128, and there were no effects on daily sperm production. The weight of the pituitary gland was reduced in male offspring only at all doses but not in a dose-dependent manner (maximum reduction, 42–52% at the lowest dose). TCDD also accelerated the average time to eye opening by about 1 day and decreased thymus weights (not dose-dependent) on postnatal day 44 in male but not female offspring. No urogenital tract malformations were seen in female offspring, and the time to vaginal opening was not affected. The uterine weights of mice both in and out of estrus were significantly reduced by about 34% at the highest dose only. Ovarian weights were not affected at any time (Theobald & Peterson, 1997).

In pregnant Long Evans rats, single doses of TCDD at 0.05–1 ΅g/kg bw per day on day 15 of gestation, which are associated with a variety of genital malformations in male and female pups, resulted in fetal concentrations on day 16 that were 0.51–0.17% of the maternal dose (13 pg/kg at 0.2 ΅g/kg dose) (Hurst et al., 2000b). Similar fetal concentrations of TCDD have been reported after exposure of adult animals to 10 ng/kg bw per day for 90 days before breeding. For comparison, the single dose of 0.05 ΅g/kg on day 15 of gestation resulted in a maternal body burden on day 21 of 27 ng/kg bw (Hurst et al., 1998a).

In a preliminary study, pregnant Wistar rats were treated on day 15 of gestation with a single oral dose of 3,3΄,4,4΄-TCB at 100 ΅g/kg bw (n = 17), 3,3΄,4,4΄,5-PeCB at 10 ΅g/kg bw (n = 23) or solvent (n = 20), and male pups were assessed for developmental end-points during lactation, on postnatal day 65 (puberty) and on postnatal day 140 (adult). At birth, the offspring of females given 3,3΄,4,4΄-TCB were slightly heavier than controls, the weight difference increasing to about 15% by the end of lactation; however, by postnatal days 65 and 140, the male pups were 8–12% lighter than controls. Whereas pups of dams given 3,3΄,4,4΄,5-PeCB weighed slightly less than controls at birth, no further significant effects on body weight were observed during the remainder of the experiment. 3,3΄,4,4΄-TCB treatment also induced slight increases in the weights of pup testis and epididymis and in daily sperm production, but the weight of the seminal vesicle and serum testosterone concentration (only on postnatal day 140) were reduced. The effects observed in pups of dams given 3,3΄,4,4΄,5-PeCB included a reduced ratio of anogenital distance:body length (postnatal day 22), a delay of 8 days in vaginal opening, slight reductions in ventral prostate weights and an increased number of mounts with intromission when compared with either controls or dams given 3,3΄,4,4΄-TCB. The serum testosterone concentrations on postnatal day 140 were also reduced in male pups of dams given 3,3΄,4,4΄,5-PeCB by up to 64%, but reproductive performance was not affected in either group treated with PCBs (Faqi et al., 1998).

In a study designed to determine the most sensitive effect of TCDD on the rodent male reproductive system, groups of six pregnant Holtzman rats were given a single oral dose of TCDD at 0, 12.5, 50, 200 or 800 ng/kg bw on day 15 of gestation, and the male offspring were assessed for reproductive development on postnatal days 49 and 120. TCDD had no effect on the body-weight gain of dams during gestation or on the weight of male pups during the assessment period. Maternal treatment with TCDD also had no effect on the relative testis or epididymal weights, daily sperm production, cauda epididymal sperm reserve or serum luteinizing hormone, follicle-stimulating hormone or testosterone concentrations of male offspring. At postnatal day 120, the maternal dose of 800 ng/kg bw resulted in TCDD concentrations of 24 pg/g of adipose tissue and 0.49 pg/g of testis in male offspring. Although the anogenital distance was not affected on postnatal day 49, it was significantly reduced at doses > 50 ng/kg bw on postnatal day 120 (maximum decrease, about 18%). In addition, on postnatal day 120 the weight of the ventral prostate was significantly reduced in pups of dams at 200 and 800 ng/kg bw, by 27% and 39%, respectively. Semi-quantitative reverse transcription polymerase chain analysis indicated that the levels of androgen receptor mRNA in this organ were significantly reduced in all treated groups on postnatal day 49 but not on postnatal day 120. Administration of TCDD at any dose resulted in a dose-dependent increase in 5alpha-reductase type 2 mRNA and a decrease in androgen receptor mRNA in the ventral prostate of rats killed on day 49 but not in those killed on day 120, with no adverse sequelae at the lowest dose of 12.5 ng/kg bw. Although the authors postulated that the decreased size of the ventral prostate was due to the decrease in androgen receptor mRNA, previous results (Gray et al., 1995) with Long Evans rats indicated that maternal exposure to TCDD at doses up to 0.8 ΅g/kg bw did not affect the number of androgen receptors in the ventral prostate of male offspring (Ohsako et al., 2001).

2.2.6 Special studies

(a) Neurotoxicity

There is currently limited experimental evidence for an Ah receptor-based mechanism of action for dioxin-induced neurotoxicity. Central nervous system functioning and neurobehaviour have been assessed in some animal models.

Pregnant Wistar rats were treated by subcutaneous injection with a loading and maintenance dosing regime designed to approximate human perinatal exposure. After an initial dose of TCDD at 0, 1 or 0.3 ΅g/kg bw on day 19 of gestation, the animals were given a weekly maintenance dose of 0, 0.4 or 0.12 ΅g/kg bw throughout lactation, and the weaned pups were given the same doses until postnatal week 11. Offspring of dams at the highest dose had significantly reduced body weight (by about 15%) throughout lactation, but not after postnatal day 31. After a slight initial decrease in body weight on postnatal day 7, the body weights of pups of dams on the low-dose regimen were comparable to those of controls. Greater percentages of the TCDD-treated offspring than controls acquired a righting reflex on postnatal day 5 or 6, but they showed a delayed ability to remain on a rotating rod. The activity of 3-month-old female offspring was significantly reduced after treatment with TCDD in comparison with controls after an amphetamine challenge. The authors concluded that TCDD may induce neurobehavioural changes (Thiel et al., 1994).

Groups of three viable rhesus offspring, born after groups of eight maternal animals had ingested a diet containing TCDD at 5 ng/kg bw for 16.2 months, were tested for aspects of peer-group interactive behaviour at 8.6 months of age. Monkeys treated with TCDD showed an increased overall level of behavioural arousal, including increased self-directed behaviour, when compared with controls. No significant relationship was observed between any of the behavioural parameters and the concentration of TCDD in the body fat of the infants. Similar effects were not seen when a second cohort of offspring obtained after the maternal animals had been on control diet for 16 months were tested. At that time, the mean concentration of TCDD in the adipose tissue of the offspring was about 50% less (188 pg/g) than that of the first cohort (Schantz et al., 1992).

A further group of infant rhesus monkeys (one female, four males) was obtained after the maternal animals had been on the TCDD-containing diet for 36.3 months. The total TCDD consumption of the monkeys was estimated to have been 880 ng, and the concentrations of TCDD in mesenteric fat obtained when the offspring were 5 months of age were similar to that in the first cohort (320 pg/g). Cognitive testing of the monkeys in both groups was conducted when they were about 14 months of age. Monkeys exposed perinatally to TCDD took a significantly longer time to learn the first reversal of a shape discrimination–reversal problem than controls (47 versus 27 trials). No significant effect of TCDD was seen for subsequent reversals (up to eight). In addition, no significant effect was seen on learning of spatial or colour reversal problems or in the ability to respond to a delayed spatial alternation test. The performance of offspring on a shape discrimination–reversal problem was not correlated to the concentration of TCDD in their adipose tissue. The authors reported that similar cognitive effects were observed in rhesus infants that had been exposed during development to lead (Schantz & Bowman, 1989).

(b) Immunotoxicity

TCDD-induced immunosuppression has been observed in several experimental animal species, including rodents, guinea-pigs, rabbits and non-human primates (reviewed by Kerkvliet & Burleson, 1994; Holsapple, 1995). Several tissues and cells of the immune system are targeted by TCDD (Kerkvliet & Burleson, 1994).

TCDD-induced thymic atrophy has been observed after perinatal exposure in all species examined (Holladay et al., 1991; Blaylock et al., 1992; Gehrs & Smialowicz, 1997; Gehrs et al., 1997; Gehrs & Smialowicz, 1999). TCDD-induced thymic atrophy is probably of less clinical significance in adult animals, as no correlation has been established between effects on the thymus and functional immune suppression. The reported acute ED50 values for thymic atrophy in adult animals vary considerably among species: 26 ΅g/kg bw in Sprague-Dawley rats, 0.8 ΅g/kg bw in Hartley guinea-pigs, 280 ΅g/kg bw in C57Bl/6 mice and 48 ΅g/kg bw in Syrian hamsters (Hanberg et al., 1989). Several hypotheses have been proposed to explain the mechanism by which TCDD induces thymic atrophy. Apoptosis induced by TCDD and changes in molecules encoded by the major histocompatibiity complex (MHC) may contribute to the thymic atrophy seen after exposure to TCDD (McConkey et al., 1988; Gerschenson & Rotello, 1992). Rhile et al. (1996) investigated the role of Fas (CD95), an important molecule involved in the induction of apoptosis, on TCDD-mediated immunotoxicity in mice bearing a homozygous lpr mutation, which leads to failure of expression of Fas. TCDD administered orally at 0.0, 0.1, 1 or 5 ΅g/kg bw for 11 days was less toxic to thymocytes from C57Bl/6 lp/lpr mice (Ah-responsive, Fas–) than to C57Bl/6 +/+ mice (Ah-responsive, Fas+). Similar results were obtained when T-cell responsiveness was tested with an antigenic challenge with conalbumin, suggesting that TCDD acts at a time when T cells are in the process of differentiating in response to antigenic challenge, as TCDD had no effect on naïve or resting T cells that had not been challenged with an antigenic stimulus. When mice that differed only at the MHC locus were compared for immunotoxic effects of TCDD, it was observed that B10.D2 (Ah-responsive, H-2d) mice were more sensitive to TCDD-induced thymic atrophy and altered peripheral T-cell function than B10 mice (Ah-responsive, H-2b). Thymic atrophy in all TCDD-sensitive strains was accompanied by depletion in T cell subsets (CD4+, CD4+ CD8+, CD4– CD8– and CD8+). On the basis of these results, the authors suggested that the observed immunotoxic effects of TCDD may be mediated partly through the presence of the Fas molecule in activated T cells and that MHC phenotype may also play an important role (Rhile et al. 1996).

TCDD affects both cellular and humoral immunity. Its effects on T-cell function are characterized by changes in several end-points, including delayed-type hypersensitivity (DTH) responses, rejection of skin allografts, generation of cytotoxic T lymphocytes and lymphoproliferative responses of lymphocytes to mitogens and specific antigens in vitro. The effects on T-cell function occur at concentrations three orders of magnitude lower than those that affect thymus cellularity. Species differ in their sensitivity to TCDD, guinea-pigs being more sensitive than rodents. For example, the DTH response to tuberculin was decreased in guinea-pigs after eight weekly doses of 40 ng/kg bw TCDD, whereas in rats the DTH response to tuberculin was unaffected by six weekly doses of 5 ΅g/kg bw (Vos et al., 1973).

Differences in DTH response have also been observed in the same species challenged with different antigens. In C57Bl/6 mice given TCDD as four weekly intraperitoneal doses of 0, 0.4, 4 or 40 ΅g/kg bw, the DTH response to sheep red blood cells (SRBC), footpad swelling, and to oxazolone, ear thickness, was measured. TCDD at doses of 4 and 40 ΅g/kg bw suppressed the DTH response to oxazolone, whereas the DTH response to SRBC was not significantly affected by any of the doses tested (Clark et al., 1981).

TCDD-induced effects on the cytotoxic T lymphocyte response of spleen cells have been found less consistently. Clark et al. (1981) treated C57Bl/6 mice with TCDD as four weekly intraperitoneal doses of 0, 0.4, 4 or 40 ΅g/kg bw and examined the generation of cytotoxic T lymphocytes. Generation was inhibited at all doses. Mechanistic studies by the same group indicated that generation of allospecific cytotoxic T cells in C57Bl/6 mice was significantly impaired after exposure to doses as low as 4 ng/kg bw (Clark et al., 1981). On the basis of mechanistic studies in vitro, the authors concluded that the cellular basis for suppression of the cytotoxic T lymphocyte response was induction of T-suppressor cells in the thymus which act selectively against the cytotoxic T lymphocyte response (Clark et al., 1983). Dooley et al. (1990) found, however, that the number of T-suppressor cells in splenocytes of B6C3F1 mice treated with TCDD by gavage at 1 ΅g/kg bw for 5 days was not increased. The mechanism by which TCDD affects the cytotoxic T lymphocyte response therefore remains to be elucidated.

Immunosuppression, indicated by suppression of the anti-SRBC humoral immune response, has been observed by several investigators (reviewed by Kerkvliet & Burleson, 1994). The dose–response relationships for the immunosuppressive effects of TCDD, 2,3,4,7,8-PeCDF, 1,2,3,7,9-PeCDF, 2,3,7,8-TCDF and 1.3.6.8-TCDF on the splenic plaque-forming cell response to SRBC in C57Bl/6 mice, measured as the ED50, were 2.4 nmol/kg bw for TCDD, 3 nmol/kg bw for 2,3,4,7,8-PeCDF, 14 nmol/kg bw for 2,3,7,8-TCDF, 710 nmol/kg bw for 1,2,3,7,9-PeCDF and 36 000 nmol/kg bw for 1,3,6,8-TCDF (Davis & Safe, 1988). Mice of different strains had different susceptibility to aryl hydrocarbon hydroxylase induction and different levels and/or affinity for a specific cytosolic binding protein; they also differed in their response to SRBC. Specifically, TCDD at a dose as low as 1.2 ΅g/kg bw significantly inhibited the plaque-forming cell response to SRBC in C57Bl/6 and C3H/HeN mice, which are representative of more susceptible strains, while a dose of at least 6 ΅g/kg bw was required for partial suppression in DBA/2 and AKR strains. Exposure for up to 8 weeks did not increase the sensitivity of DBA/2 mice to treatment (Vecchi et al., 1983). The plaque-forming cell and immunoglobulin M responses of C57Bl/6 and DBA/2 mice to the T cell-independent antigen trinitrophenyl-lipopolysaccharide also varied with different congeners, with ED50 values in C57Bl/6 mice of 2.8 and 1.6 ΅g/kg bw for TCDD, 11 and 14 ΅g/kg bw for 3,3΄,4,4΄,5-PeCB and 25 and 20 ΅g/kg bw for 3,3΄,4,4΄,5,5΄-HxCB, and values in the less Ah-responsive DBA/2 mice for the same congeners of 8.5 and 10 ΅g/kg bw, 61 and 69 ΅g/kg bw and 73 and 71 ΅g/kg bw, respectively (Harper et al., 1994).

Differences among species in the primary response to antigens have also been observed. B6C3F1 mice and Fischer 344 rats were given a single intraperitoneal injection of TCDD in corn oil at a dose of 1, 3, 10 or 30 ΅g/kg bw on 7 days before immunization with SRBC for induction of a humoral response. The plaque-forming cell response in mice was significantly suppressed, with an ED50 value of 0.7 ΅g/kg bw, while the response in rats was unaffected at doses up to 30 ΅g/kg bw. The apparent inability of TCDD to suppress humoral immunity in rats was unrelated to induction of hepatic CYP 1A1 and CYP 1A2. Furthermore, in rats but not in mice, the splenic CD4– CD8+ T-cell sub-population was reduced in a dose-related manner, and this reduction was accompanied by a dose-related increase in immunoglobulin M+ splenocytes (Smialowicz et al., 1994). In a similar experiment with the same doses of TCDD, the response of B6C3F1 mice to the T cell-independent, B cell-dependent antigen trinitrophenyl-lipopolysaccharide was reduced at 10 and 30 ΅g/kg bw dose, and a similar reduction was seen in rats at 30 ΅g/kg bw, with no significant changes in splenic lymphocyte subsets. The mechanism responsible for the observed differences between the two species in their response to T cell-dependent and T cell-independent antigens remains to be elucidated (Smialowicz et al., 1996).

Differences in the response to various antigens after exposure to TCDD have also been observed within the same species. Antibody responses to three antigens, SRBC, DNP-Ficoll and trinitrophenyl-lipopolysaccharide,were compared in C57Bl/6 mice after treatment with 1,2,3,4,6,7,8-HpCDD. The median inhibitory doses were 53, 130 and 520 ΅g/kg for the three antigens, respectively, indicating that higher doses of TCDD are required to suppress the T cell-independent, B cell-dependent humoral immune response (Kerkvliet & Brauner, 1987).

The effects of TCDD on host resistance to various infectious agents, including bacteria, viruses and parasites, has been well documented. Four-week-old male C57B1/6J (J67) mice received TCDD by gavage at a dose of 0.5–20 ΅g/kg bw once a week for 4 weeks and were challenged 2 days after the fourth dose with either Salmonella bern or Herpesvirus suis. Increased mortality in response to S. bern was seen at 1 ΅g/kg bw and a shorter latency before death at 5 ΅g/kg bw. None of the doses compromised the ability of mice to combat infection by H. virus suis (Thigpen et al., 1975). Compromised resistance to Salmonella was also shown in mice given feed containing TCDD at 50 or 100 ng/kg for 8 weeks (Hinsdill et al., 1980).

Increased parasitic infestation has been reported after administration of TCDD. A single dose of 5 or 10 ΅g/kg bw administered by gavage increased the susceptibility of 6–8-week-old female B6C3F1 mice to Plasmodium yoelii 17 XNL, which causes malaria (Tucker et al., 1986). A dose of 10 or 30 ΅g/kg bw administered intraperitoneally 7 days before infection of B6C3F1 mice with Trichinella spiralis resulted in delayed onset of parasite elimination, while a dose of 1 ΅g/kg bw suppressed the proliferative response of splenocyte and mesenteric lymph node cells stimulated with T. spiralis antigen (Luebke et al., 1994). In contrast, the same doses of TCDD given to adult Fischer 344 rats 7 days before infection with T. spiralis had no effect on the rate of elimination of the parasite or the number of encysted larvae in muscle, and the dose of 30 ΅g/kg bw enhanced the proliferative response of lymphocytes after stimulation with the parasite antigen (Luebke et al., 1995).

The ability of hosts to combat viral infection was compromised at doses of TCDD lower than those that affected resistance to bacteria and parasites. Mice given an intraperitoneal injection of TCDD at 0.04, 0.4 or 4 ΅g/kg bw once a week for 4 weeks and challenged 7–22 days later with Herpes simplex type II strain 33 virus had a significantly enhanced death rate from viral infection (Clark et al., 1981). In female B6C3F1 mice, 6–8 weeks of age, given TCDD at 10, 1 or 0.1 ΅g/kg bw intraperitoneally and challenged with A/Taiwan/1/64 (H2N2) virus 7–10 days later, the LOEL for decreased resistance to virus was 0.1 ΅g/kg bw, the lowest dose tested (House et al., 1990).

An even lower LOEL of 0.01 ΅g/kg bw was seen for enhanced mortality from influenza due to A/Hong Kong/8/68 (H3N2) virus. Eight-week-old female B6C3F1 mice were given a single dose of TCDD at 0.1, 0.05 or 0.01 ΅g/kg bw in corn oil by gavage. Seven days later, the mice were infected intranasally with a concentration of Influenza A/Hong Kong/8/68 (H3N2) virus calculated to result in 30% mortality in mice untreated with TCDD, and the animals were observed for 22 days. The mortality rates of mice given TCDD was statistically significantly higher than that of controls, but the rate at doses of 0.001 and 0.005 ΅g/kg bw was not significantly different from that of controsl. TCDD did not alter the replication or clearance of the virus in these mice (Burleson et al., 1996).

Young animals are reported to be highly sensitive to prenatal or neonatal exposure to TCDD. Pregnant B6 or B6C3F1 mice were given TCDD oraly at a dose of 1, 2, 5 or 15 ΅g/kg bw on days –7, 0, +7 and +14 relative to parturition. The LOEL for an increased incidence of PYB6 tumours was 1 ΅g/kg bw, the lowest dose tested. The LOAEL for increased allograft rejection time was 2 ΅g/kg bw and that for decreased mortality from L. monocytogenes infection, T-cell blastogenesis, thymus and spleen weights, bone-marrow cellularity and bone-marrow colony forming units was 5 ΅g/kg bw. The LOEL for lipopolysaccharide blastogenesis and anti-SRBC serum titres was > 15 ΅g/kg bw (Vos & Moore, 1974; Luster et al., 1980).

Pregnant Swiss Webster mice were fed diets containing TCDD at 1, 2.5 or 5 ΅g/kg for 7 weeks before and after parturition. The LOEL for increased mortality of offspring due to endotoxin was the lowest dose, that for decreased thymus weight was 2.5 ΅g/kg, and that for decreased plaque-forming cell response to SRBC and DTH responses was 5 ΅g/kg. The LOEL for increased mortality due to L. monocytogenes, T and B cell blastogenesis and anti-SRBC serum titres was > 5 ΅g/kg of diet (Thomas & Hinsdill, 1979).

When pregnant Fischer 344 rats were given TCDD at 1 or 5 ΅g/kg bw per day orally on days –3, 0, +7 and +14 relative to parturition, the LOEL for increased allograft rejection time, for decreased T cell blastogenesis, DTH response and body and thymus weight and for increased L. monocytogenes-induced mortality was 5 ΅g/kg bw per day (Vos & Moore, 1974; Faith & Moore, 1977).

A single dose of TCDD at 0.1, 0.3, 1 or 3 ΅g/kg bw given by gavage to pregnant Fischer 344 rats on day 14 of gestation resulted in suppression of the DTH response to bovine serum albumin in both male and female pups. The suppression persisted for up to 19 months in the male offspring of dams given 3 ΅g/kg bw. Exposure during both gestation and lactation was required for suppression of the DTH response (Gehrs et al., 1997; Gehrs & Smialowicz, 1999).

Further experiments with TCDD at 0, 0.1, 0.3 or 1 ΅g/kg bw indicated that doses as low as 0.1 ΅g/kg bw given to dams on day 14 of gestation could suppress the DTH response in males up to 14 months of age, while a maternal dose of 0.3 ΅g/kg bw was necessary to cause suppression that continued to 14 months in female offspring. The DTH response to a second antigen, keyhole lympet haemocyanin, was also reduced in the 4-month-old male offspring of dams dosed at 3 ΅g/kg bw on day 14 (Gehrs & Smialowicz, 1999). Thus, doses of TCDD as low as 0.1 ΅g/kg bw administered orally to dams on day 14 of gestation caused significant, long-lasting effects on the DTH response of offspring, males being more sensitive to treatment than females.

Groups of six male Wistar rats were maintained on diets containing TCDD at 0, 0.34 or 5 ΅g/kg for 30 or 180 days. The TCDD-containing diets were given on 5 days/week and the control diet on 2 days/week. The total intake of TCDD was about 3 ΅g/kg bw or 100 ng/kg bw per day for 30 days or 17 ng/kg bw per day for 180 days. All rats were killed 3 days after the last feeding. No significant effects were seen on body-weight gain or relative spleen or liver weight, while the relative thymus weight was reduced by about 27% in rats treated for 30 days. The mitogenic proliferative response of spleen cells to lipopolysaccharide in vitro was enhanced in the group treated for 30 days to lipopolysaccharide, and the response to concanavalin A stimulation was slightly decreased in the group treated for 180 days, in comparison with controls. Production of IL-1 by spleen macrophages in vitro was significantly decreased in both groups, by 38% in those treated for 30 days and by 52% in those treated for 180 days. Stimulation of IL-2 production by spleen cells was also reduced, but only in the group treated for 180 days. Both groups showed decreased conconavalin A-induced expression of IL-2 receptors by splenic T-cells in vitro (Badesha et al., 1995).

In summary, immunotoxic effects of TCDD have been observed in several species at multiple targets in the immune system. The severity of TCDD-induced immunotoxic effects varies among species and depends largely on the end-point investigated. On the basis of the above review, the LOEL for suppressed DTH responses in offspring of perinatally exposed Fischer 344 rats was 0.1 ΅g/kg bw, and the LOEL for increased susceptibility of adult B6C3F1 mice to H2N2 virus was 0.01 ΅g/kg bw, with a NOEL of 0.005 ΅g/kg bw.

(c) Effects on the thyroid

The effects on the thyroid observed in experimental animals after exposure to dioxins or coplanar chemicals usually involve decreases in free and total T4 in serum, with no compensatory effects on TSH or T3 (Brucker-Davis, 1998). Dose-dependent decreases in plasma T4 concentration were observed in groups of eight female Sprague-Dawley rats given diets supplemented with TCDD at a concentration of 0, 0.2, 0.4, 0.7, 5 or 20 ΅g/kg for 13 weeks, with estimated daily intakes of 0, 14, 26, 47, 320 and 1000 ng/kg bw per day. The concentration of total T4 in plasma was significantly reduced at doses > 47 ng/kg bw per day (van Birgelen et al., 1995b). Similar effects were observed in groups of eight female rats of the same strain given diets containing 3,3΄,4,4΄,5-PeCB at a concentration of 0, 7, 50 or 180 ΅g/kg for 13 weeks, providing daily intakes of 0, 0.47, 3.2 and 10 ΅g/kg bw per day. The lowest dose of 3,3΄,4,4΄,5-PeCB that induced a significant decrease in free or total T4 in plasma was 3.2 ΅g/kg bw per day. The LOEL was 0.047 ΅g/kg bw per day (van Birgelen et al., 1995a).

Thyroid hormone status was assessed in rats from a short-term assay for tumour promotion. After initiation with N-nitrosodiethylamine at 70 days of age, female Sprague-Dawley rats were treated by gavage every 2 weeks for 30 weeks with TCDD at doses designed to deliver 0, 0.1, 0.35, 1, 3.5, 11, 36 or 120 ng/kg bw per day. The rats were necropsied 1 week after the last dose of TCDD, and serum samples were analysed for T3, T4 and TSH. The concentration of T4 was significantly reduced at doses of TCDD > 11 ng/kg bw per day in the initiated rats and > 36 ng/kg bw per day in the uninitiated rats; the maximum reduction was by 42%. While there was no effect on T3 concentration, that of TSH in serum was increased approximately 2.5-fold in uninitiated rats at the highest dose when compared with controls; 3.3 ng/ml and 1.3 ng/ml, respectively. Histological changes in the thyroid, including diffuse follicular hyperplasia, were also seen in TCDD-treated rats, and, at doses > 3.5 ng/kg bw per day, the ratio of parenchymal to follicular area was significantly increased. Hepatic CYP 1A1 and UGT1 mRNA levels were increased at doses > 0.35 ng/kg bw per day and 3.5 ng/kg bw per day, respectively (Sewall et al., 1995).

Whereas other related polyhalogenated aromatic hydrocarbons, including some coplanar PCBs, can interact directly with the thyroid gland and/or thyroid hormone transport mechanisms, dioxins and furans appear to function primarily through hormone metabolism (Brouwer et al., 1998). Significant positive correlations have been observed between TCDD-induced decreases in plasma T4 concentration and concomitant induction of hepatic UGT1, indicating that TCDD functions predominantly at an extrathyroidal level (Schuur et al., 1997). Further evidence for a mechanism of action that does not directly involve the thyroid gland was obtained in the long-term bioassay in which male and female Sprague-Dawley rats were given TCDD at doses of 0.001–0.1 ΅g/kg/bw per day for up to 2 years, with no increase in the incidence of non-neoplastic lesions in the thyroid, parathyroid, adrenal or pituitary glands (Kociba et al., 1978b).

On the basis of TEFs (in parentheses), a mixture comprising TCDD (1), 1,2,3,7,8-PeCDD (3.3), 2,3,4,7,8-PeCDF (17), 3,3΄,4,4΄,5-PeCB (61), 2,3΄,4,4΄,5-PeCB (12 800) and 2,3,3΄,4,4΄,5-HxCB (1888) was more potent than TCDD in reducing the total plasma concentration of T4 in groups of 10–12 female Sprague-Dawley rats after subcutaneous injection of a toxic equivalent of 1 ΅g/kg bw per week (van der Plas et al., 2001). TCDD at 1 ΅g/kg bw per week reduced the total plasma T4 concentration by 38%, while a toxic equivalent of the mixture reduced it by 54%. Indications of reduced T4 protein binding (decreased ratio of total:free T4) suggested competitive displacement of T4 from the major rodent thyroid hormone transport protein, transthyretin, possibly due to the presence of hydroxylated PCB metabolites from the mixture. Various hydroxylated metabolites of PCBs have been shown to competitively displace T4 binding to human transthyretin but not to T4-binding globulin (Lans et al., 1994).

Groups of six or seven male Sprague-Dawley rats treated by gavage with TCDD at a dose of 0, 0.02, 0.23, 1.2, 3.5, 7 or 12 ΅g/kg bw per week for 10 weeks showed dose-dependent decreases in total serum T4 at doses > 0.23 ΅g/kg bw per week. Four of seven rats at the highest dose died, and decreased body-weight gain was seen at the three higher doses. After a 6-week recovery period, at which time it was estimated that about 75% of the administered TCDD would have been eliminated, the total T4 concentration in serum continued to be reduced in groups with a cumulative total dose of TCDD > 35 ΅g/kg bw (Li & Rozman, 1995).

Groups of 10–14 pregnant Sprague-Dawley rats were given 3,3΄,4,4΄-TCB at 2 or 8 mg/kg bw per day, 3,3΄,4,4΄,5-PeCB at 0.25 or 1 ΅g/kg bw per day, TCDD at 0.025 or 0.1 ΅g/kg bw per day or solvent orally on days 10–16 of gestation. Treatment had no effect on the weight gain of dams during gestation, on litter size or on pup survival or growth during lactation. When selected pups were necropsied on postnatal day 21, the absolute thymus weights were decreased only at the higher doses of 3,3΄,4,4΄-TCB and TCDD, by 15% and 23%, respectively, while the liver weights were significantly increased at both doses of 3,3΄,4,4΄,5-PeCB, by 13% and 17%, respectively. Perinatal treatment had no effect on the thyroid hormone status (T3, T4, TSH) of male pups, and only the serum T4 concentration was decreased in female pups, by 15–20% at the higher doses of 3,3΄,4,4΄-TCB and TCDD. Hepatic UGT1 and EROD activity were increased by all treatments except the lower dose of 3,3΄,4,4΄-TCB; UGT1 activity was increased by up to threefold and EROD activity by up to 24-fold with the higher doses of 3,3΄,4,4΄,5-PeCB and TCDD. No sex-specific differences were seen with respect to enzyme induction (Seo et al., 1995).

(d) Vitamin A status

Characteristic symptoms of vitamin A deficiency have been observed in experimental animals and wildlife species after exposure to TCDD and various coplanar chemicals. In rats given single oral doses of TCDD at 1–10 ΅g/kg bw, the hepatic stores of retinol were reduced and those in serum, kidney, urine and faeces were increased (excretion). The mechanism of action appears to be persistent impairment of the ability of liver stellate cells to store vitamin A due to inhibition of lecithin:retinol acyltransferase. Treatment of male Sprague-Dawley rats with a single oral dose of 10 ΅g/kg bw resulted in a significant decrease in the activity of this enzyme in the non-parenchymal liver cell fraction 7 days later (Nelson et al., 1996). The initial kinetic effects after exposure to TCDD involve increased mobilization of vitamin A from hepatic storage sites, probably by retinal ester hydrolysis (Kelley et al., 2000).

Guinea-pigs, rats, Ah-responsive and Ah-unresponsive strains of mice (C57Bl/6 and D/A/2, respectively) and hamsters were given a single intraperitoneal dose of TCDD representing 9–67% of their respective LD50 value (0.5–400 ΅g/kg bw) and killed at various times up to 112 days after treatment. Hepatic and pulmonary vitamin A stores were decreased in all species except C57Bl/6 mice (liver only), the degree of change being related to obvious signs of toxicity (lethality, decreased growth rate, liver hypertrophy, thymic involution) (Hawkinston et al., 1979).

Male Sprague-Dawley rats were given a single oral dose of TCDD at 0, 0.1, 1, 10, 30, 100 or 300 nmol/kg bw (approximately 0, 0.032, 0.32, 3.2, 9.6, 32 and 96 ΅g/kg bw), and three rats at each dose were killed 12 days later for analysis of retinol, retinal palmitate, retinal palmitate hydrolase and acyl coenzyme A:retinol acyltrans-ferase activity in kidney and liver. Additional control groups consisted of rats maintained on a vitamin A-free diet for 12 or 26 days. The relative liver weights were significantly increased in a dose-dependent manner in rats at doses > 0.32 ΅g/kg bw, except those at the highest dose, which had an approximately 50% loss of body weight. The relative kidney weights were significantly increased only at the lowest dose. Rats fed the vitamin A-free diet for 12 days had increased relative liver weight similar to that observed in rats given the three lower doses of TCDD (10–22%). The concentrations of hepatic retinal palmitate were decreased at all doses of TCDD when compared with controls, with a maximum reduction of 75% in the group at the highest dose. The concentrations of hepatic retinal palmitate levels were also reduced in the rats fed the vitamin A-free diet for 12 days to an extent similar to that of rats given the three lower doses of TCDD (about 40%). The concentrations of retinol and retinal palmitate in kidney were increased at doses of TCDD of 3.2, 9.6 and 32 ΅g/kg bw and in the rats fed the vitamin A-free diet for 26 days. The activity of acyl coenzyme A:retinol acyltransferase in kidney was increased only at 9.6 and 32 ΅g/kg bw and in rats fed the vitamin A-free diet for 26 days (Jurek et al., 1990).

Groups of six male and six female Sprague-Dawley rats were fed diets containing various concentrations of dioxins and furans for 13 weeks, and hepatic retinol concentrations were determined at sacrifice. The estimated median effective dietary concentrations of TCDD for reducing hepatic retinol were 0.5 ΅g/kg of diet for females and 0.8 ΅g/kg of diet for males, equivalent to 50 and 80 ng/kg bw per day, respectively, while the values for the next most potent congener, 1,2,3,7,8-PeCDD, were 0.9 ΅g/kg and 1.2 ΅g/kg of diet, respectively. The ranking of potency was TCDD > 1,2,3,7,8-PeCDD > 2,3,4,7,8-PeCDF > 1,2,3,6,7,8-HxCDF > 1,2,3,7,8-PeCDF > OCDF > OCDD > 1,2,3,4,8-PeCDF. The estimated relative potency factors for hepatic vitamin A depletion corresponded to the values for indices of toxicity in short-term tests (thymic atrophy, body-weight reduction, liver lesions). Dietary intake of a variety of chlorinated dioxins and furans for 13 weeks thus reduced hepatic vitamin A concentrations in a dose-dependent manner (Fattore et al., 2000).

As retinol is transported in blood bound to a complex of retinol binding protein and transthyretin, destabilization of this complex could reduce the binding ability of retinol, with subsequent elimination by renal glomerular filtration. In the study of van der Plas et al. (2001), a mixture of coplanar PCBs caused a decrease in hepatic retinal palmitate concentration similar to that induced by an equivalent dose of TCDD alone. However, only the mixture reduced plasma retinol concentrations, presumably by the generation of hydroxylated PCB metabolites and effects on retinol binding protein–transthyretin binding.

When groups of six male weanling Sprague-Dawley rats were given a single intraperitoneal injection of 3,3΄,4,4΄-TCB or 3,3΄,4,4΄,5-PeCB at a dose of 150 ΅mol/kg and killed 7 days later, significant decreases in the concentrations of hepatic retinyl palmitate and serum retinol were observed. Both PCBs significantly increased the relative liver weights and decreased the relative thymus weights, while 3,3΄,4,4΄,5-PeCB induced a 2.7-fold increase in hepatic lipid accumulation, when compared with controls. 3,3΄,4,4΄,5-PeCB-treated rats also lost approximately 20 g of body weight. 3,3΄,4,4΄-TCB treatment caused a significant increase (approximately 10-fold) in the renal stores of retinyl palmitate over that of controls, and both PCBs decreased the serum retinol binding protein concentration (Chen et al., 1992a). Various hydroxylated metabolites of coplanar PCBs have been shown to bind with high affinity to transthyretin, resulting in inhibition of T4 binding to this protein and disruption of the transthyretin–retinol binding protein complex (Brouwer et al., 1998).

2.3 Observations in humans

The epidemiological studies on dioxins include studies of the exposure of workers producing phenoxy herbicide and chlorophenols, studies of the population exposed in the industrial accident in Seveso, Italy, studies of persons exposed during application of herbicides and particularly cohorts of personnel in the US Air Force in Viet Nam, commercial applicators and case–control studies of exposure of communities.

The most informative epidemiological studies are those of the population of Seveso who were accidentally exposed to dioxins in 1976, those of workers producing chlorophenols and chlorophenoxy herbicides contaminated with dioxins and US Air Force applicators. Considerable effort has been made to characterize the exposure of these populations to TCDD, which was 10–1000 times higher than that of the general population. The populations included in these studies are shown in Table 5.

Table 5. Populations included in the most informative epidemiological studies of the effects of exposure to dioxin

Country (reference)

Number of subjects

Concentration of TCDD in blood

Outcomes examined

Seveso, Italy, industrial accident (Bertazzi et al., 1993; Mocarelli et al., 1996; Bertazzi et al., 1997; Landi et al., 1997)

Population residing in Seveso area; zone A (most heavily contaminated), 750 persons; zone B, 5000 persons; zone R, 30 000 persons

Geometric mean in1996: zone A, 7 individuals, 53 pg/g; zone B, 11 pg/g; non-A, non-B, non-reference zone, 4.9 pg/g

Mortality, cancer incidence, morbidity, biochemical parameters in adults and children

Germany, accident at BASF plant (Zober et al., 1990; Ott & Zober, 1996)

247 (243 men, 4 women) directly involved in accident or clean-up

Geometric mean in 1988–92, 15 pg/g; very high concentrations in workers with chloracne

Mortality, morbidity, biochemical parameters

Germany, other plants, including Boehrin-ger (Manz et al., 1991; Nagel et al., 1994; Flesch-Janys et al., 1995; Becher et al., 1996)

2479 male workers employed in four German plants; Boehringer cohort included women

Boehringer plant, 1985–94: mean, 140 pg/g; lower in another plant; background concentraions in remaining two plants

Mortality

Netherlands, plants (Bueno de Mesquita et al., 1993; Hooiveld et al., 1998)

2074 men employed in two plants

Mean, 1993: 53 pg/g in one plant; background concentration in second

Mortality

United States, plants (Fingerhut et al., 1991; Steenland et al., 1999; Egeland et al., 1994; Calvert et al., 1999)

5172 men in 12 herbicide production plants

Average in 1987: 230 pg/g lipid

Mortality in full cohort; morbidity and biochemical parameters for small subsample

IARC, multi-country study (Saracci et al., 1991; Kogevinas et al., 1997; Vena et al., 1998)

21 863 male and female workers employed in 36 plants, 12 countries; includes all above-mentioned cohorts except BASF (Zober et al., 1990)

3–390 pg/g (574 measurements in 10 cohorts, 7 countries)

Mortality, cancer incidence

Viet Nam, US Army herbicide applicators (Michalek et al., 1990; Henriksen et al., 1996, 1997; Michalek et al., 1998)

1261 men

1984–85: mean, 46 pg/g; geometric mean, 16 pg/g

Mortality, morbidity, biochemical parameters

2.3.1 Cancer

This section is based largely on a review of the carcinogenicity of dioxins by a working group convened by the IARC (1997) and a WHO evaluation for a tolerable daily intake conducted in 1998 (Kogevinas, 2000).

(a) Cancer incidence and mortality in the population of Seveso

The incidence of and mortality from cancer were investigated in the population of Seveso that was exposed during the industrial accident in 1976. The contaminated area was subdivided into three zones (zone A, zone B and zone R), according to the average concentration of TCDD measured in soil samples. The concentrations in serum in 1976 in 19 non-randomly selected persons from zone A ranged from 830 pg/g to 56 000 pg/g, and the geometric mean serum concentrations measured 20 years after the accident (Landi et al., 1997) were 53 pg/g in seven randomly selected residents of zone A (the most heavily contaminated zone), 11 pg/g in 55 persons in zone B (the second most heavily contaminated zone) and 4.9 pg/g in 59 persons in the uncontaminated area (zone non-ABR). The populations of 11 municipalities surrounding the contaminated area were used as a reference (zone R).

The mortality rate during 1976–91 and cancer incidence rates for the period 1977–86 have been reported (Bertazzi et al., 1989, 1993, 1997), and follow-up of the exposed population for deaths was extended to 1996 (Bertazzi et al., 2001). The rate of death from cancer in general was not increased over the period of observation. However, 15 years after the accident, the mortality rate from cancer was increased in men among the 804 inhabitants of zone A and the 5941 inhabitants of zone B, with a rate ratio of 1.3 and a 95% confidence interval (CI) of 1.0–1.7. Increases were also seen in the rate of death from rectal cancer (rate ratio, 2.4; 95% CI, 1.2–4.6) and lung cancer (rate ratio, 1.3; 95% CI, 1.0–1.7), but with no definite pattern of latency. Lymphohaematopoietic neoplasms occurred at an excess rate in males and females (rate ratio, 1.7; 95% CI, 1.2–2.5). The mortality rate from Hodgkin disease was high during the first 10-year observation period (rate ratio, 4.9; 95% CI, 1.5–16), whereas the greatest increases in the rates of death from non-Hodgkin lymphoma (rate ratio, 2.8; 95% CI, 1.1–7.0) and myeloid leukaemia (rate ratio, 3.8; 95% CI, 1.2–12) were seen after 15 years. No cases of soft-tissue sarcoma were observed in zones A and B, although 0.8 was expected.

(b) Exposure in chemical plants

The study populations considered overlap in many instances. Figure 8 shows the relationships among the various publications considered relevant to this evaluation.

FIGURE 8

Figure 8. Relationships among various studies of industrial cohorts exposed to TCDD and higher chlorinated PCDDs adn PCDFs

(i) Germany, BASF plant

In an accident at a BASF plant producing trichlorophenol in Ludwigshafen, Germany, in 1953, a total of 247 employees (243 men, 4 women) were identified as having been involved directly or in subsequent clean-up, repair or maintenance activities (Zober et al., 1990; Ott & Zober, 1996). Measurements of TCDD in serum were available for 138 of these persons (Ott et al., 1993). The geometric mean concentration was 15 pg/g (range, < 1–550 pg/g). The cumulative dose of TCDD at the time of exposure was calculated from a model for each person. The mean concentrations were 38 pg/g for a subgroup with no chloracne, 420 pg/g for a subgroup with moderate chloracne and 1000 pg/g for a group with severe chloracne. The cohort was observed for deaths and cancer incidence until 1992, and the rate of mortality from cancer was found to increase with increasing exposure (Figure 9). Mortality rates from cancers of the digestive system and respiratory tract also tended to increase with increasing exposure. Similar results were obtained for cancer incidence. A joint analysis of dose of TCDD and cigarette smoking showed a relationship between the dose and the occurrence of all cancers combined among current cigarette smokers but not among non-smokers.

FIGURE 9

From Ott & Zober (1996); SMR, standardized mortality ratio

Figure 9. Rates of death from all cancers in relation to exposure to TCDD, BASF plant, Germany

(ii) Germany, other plants

Becher et al. (1996) reported on the mortality rates among 2479 male workers employed in four German plants involved in the production of phenoxy herbicides and chlorophenols or who were likely to have been in contact with these substances and with their contaminants, PCDDs (often including TCDD) and PCDFs. The study did not include the persons involved in the BASF accident. It included workers from a Boehringer-Ingelheim plant in Hamburg, for whom results have been reported in several publications (Manz et al., 1991; Nagel et al., 1994; Flesch-Janys et al., 1995, 1996). The concentrations of TCDD were measured in the blood of workers in three of the four plants (Kogevinas et al., 1997). The mortality rate was increased in the whole cohort from all cancers combined (138 deaths; standardized mortality ratio [SMR], 1.2; 95% CI, 1.0–1.4), cancer of the oral cavity and pharynx (nine deaths; SMR, 3.0; 95% CI, 1.4–5.6), lung cancer (47 deaths; SMR, 1.4; 95% CI, 1.1–1.9), lymphatic and haematopoietic neoplasms (13 deaths; SMR, 1.7; 95% CI, 0.9–2.9) and non-Hodgkin lymphoma (six deaths; SMR, 3.3; 95% CI, 1.2–7.1).

Several reports have been published on workers at a chemical plant operated by Boehringer-Ingelheim in Hamburg, Germany, that produced herbicides heavily contaminated with TCDD and other PCDDs and PCDFs (Manz et al., 1991; Nagel et al., 1994; Flesch-Janys et al., 1995). An outbreak of chloracne in 1954 led to a halt in the production of trichlorophenol and 2,4,5-trichlorphenoxyacetic acid. In 1957, production was resumed, with a new process that reduced the formation of TCDD. The vital status of all persons who had been permanent employees at the plant for at least 3 months between 1 January 1952 and 31 December 1984 (1184 men, 399 women) was investigated through to 1989 (Manz et al., 1991). The causes of death were ascertained from medical records or, when medical records were not available, from death certificates. The concentrations of TCDD were measured in samples of serum or adipose tissue from workers in various production departments, mainly after the plant had closed in 1984.

The rate of mortality from all cancers combined among men was increased in comparison with the rate for the general population (93 deaths; SMR, 1.2; 95% CI, 1.0–1.5). The increase was greatest for men who had started work at the plant before 1954, who had the heaviest exposure to TCDD on the basis of the history of the plant and subsequent serum measurements, and who had remained employed at the plant for many years. The mortality rates of female workers were further investigated by Nagel et al. (1994). The rate of death from breast cancer was higher than expected (10 cases; SMR, 2.4; 95% CI, 1.1–4.4) and increased with duration of employment.

The mortality rates of 1189 male workers were investigated through 1992 (Flesch-Janys et al., 1995, 1996). The concentrations of various PCDD and PCDF congeners were measured in adipose tissue or whole blood from 190 workers, and the values for each worker at the end of exposure were calculated on the basis of a one-compartment first-order kinetics model. The mean estimated concentration of TCDD for the whole cohort was 140 ng/kg of lipid (median, 38 ng/kg). In some departments, workers had been exposed to other carcinogens, such as dimethyl sulfate and benzene. The mortality rate from all cancers was increased in all categories of exposure to TCDD and showed a significant trend (p = 0.01) with increasing intensity of exposure (Figure 10). No data were reported on deaths from cancers at specific sites.

FIGURE 10

From Flesch-Janys et al. (1995)

Figure 10. Rates of mortality from all causes, all cancers and ischaemic heart disease (IHD) in relation to exposure to dioxins, Boehringer plaint, Germany

(iii) The Netherlands

Mortality rates among workers employed between 1955 and 1986 in the synthesis and formulation of phenoxy herbicides and chlorophenols in The Netherlands were examined (Bueno de Mesquita et al., 1993). In one of the plants, where the main compound produced was 2,4,5-trichlorphenoxyacetic acid and its derivatives, an accident in 1963, which caused a release of PCDDs, including TCDD. In a second factory, (4-chloro-ortho-toloxy)acetic acid, 4-chloro-2-methyl phenoxypropionic acid and 2,4-D were produced. The study involved 2074 men working in manufacture at the two plants (963 exposed to phenoxy herbicides, 1111 not exposed); in addition, 145 workers who had probably been exposed to TCDD during the industrial accident and clean-up were examined. The study was later extended to 1991 and enlarged (2298 persons, including 191 women) (Hooiveld et al., 1998). Exposure was assessed from job histories and on modelled concentrations of TCDD in serum, measured in 1993 in a subset of 31 exposed (mean concentration, 53 ng/kg of lipid; range, 1.9–190 ng/kg) and 16 unexposed (mean concentration, 8 ng/kg of lipid) persons. Fourteen persons who had been exposed during the accident in 1963 had the highest mean concentration (96 ng/kg of lipid; range, 16–190 ng/kg). In this factory, the mortality rates from all causes (139 deaths; SMR, 1.3; 95% CI, 1.1–1.5) and all cancers (51 deaths; SMR, 1.5; 95% CI, 1.1–1.9) were significantly increased. Excess numbers of deaths from cancers of the urinary bladder (four deaths; SMR, 3.7; 95% CI, 1.0–9.5) and kidney (four deaths; SMR, 4.1; 95% CI, 1.1–10) and from non-Hodgkin lymphoma (three deaths; SMR, 3.8; 95% CI, 0.8– 11) were observed. When the workers were subdivided into three categories of low, medium and high exposure according to the serum concentrations of TCDD predicted from a model, the relative risks for death from all causes, all cancers and lung cancer were significantly increased for workers with medium and high exposure and were highest for the group with the heaviest exposure (Figure 11).

FIGURE 11

From Hooiveld et al. (1998)

Figure 11. Mortality rates from all causes, all cancers and ischaemic heart disease in relation to exposure to TCDD predicted from a model, The Netherlans

(iv) United States

A study of workers in 12 plants in the USA was designed by the National Institute for Occupational Safety and Health (Fingerhut et al., 1990, 1991; Steenland et al., 1999). The cohort comprised 5172 men and included most workers in the USA likely to have been exposed to TCDD during manufacture principally of trichlorophenol and 2,4,5-trichlorphenoxyacetic acid. The average concentration of TCDD in serum from 253 workers at two plants in 1987 was 230 ng/kg of lipid, whereas that in 79 unexposed workers was 7 ng/kg. The concentration increased to 420 ng/kg in serum from 119 workers who had been exposed for more than 1 year. Extrapolation to the date at which these workers had been employed, assuming a half-life of 7.1 years, indicated a mean serum concentration at that time of 2000 ng/kg of lipid (highest concentration, 32 000 ng/kg). In the latest update, of mortality rates through 1993 (Steenland et al., 1999), the cohort was restricted to 3538 workers from eight plants for whom a detailed occupational history was available. An exposure matrix was used to estimate the degree of exposure to TCDD in the whole cohort. The SMR for all cancers combined was 1.1 (95% CI, 1.0–1.2). A statistically significant, positive, linear trend in risk for all cancers combined and for lung cancer was found with increasing exposure (Figure 12). The SMR for all cancers combined for the group with the heaviest exposure was 1.6 (95% CI, 1.2–1.8). The excess cancer risk was limited to the workers with the heaviest exposure, which was likely to have been 100–1000 times higher than that experienced by the general population and similar to the doses of TCDD used in studies in experimental animals.

FIGURE 12

From Steenland et al. (1999)

Figure 12. Mortality rates from all causes in relation to cumulative exposure to TCDD with a 15-year lag time in an exposure matrix, National Institute for Occupational Safety and Health, USA

(v) IARC multi-country study

An international cohort of workers exposed to phenoxyacetic acid herbicides and chlorophenols was set up by the IARC (Saracci et al., 1991). The cohort comprised 16 863 men and 1527 women who had been employed in production or spraying, distributed among 20 cohorts in 10 countries. The cohort was updated and expanded with the data of Fingerhut et al. (1991) and Becher et al. (1996) (Kogevinas et al., 1997). The length of follow-up differed by plant, but workers at most of the European plants were followed through 1991–92, and those in the USA through 1987. Current concentrations of TCDD were measured in 574 workers in 10 plants in seven countries ranged from 3 to 390 ng/kg of lipid. This study represents the largest cohort of TCDD-exposed workers and includes groups with heavy exposure to TCDD and cohorts with little or very little exposure. Among the workers who had been exposed to TCDD or higher chlorinated PCDDs, the mortality rate was increased for soft-tissue sarcoma (six deaths; SMR, 2.0; 95% CI, 0.8–4.4), and the rates of mortality from all cancers combined (710 deaths; SMR, 1.1; 95% CI, 1.0–1.2), non-Hodgkin lymphoma (24 deaths; SMR, 1.3; 0.9–2.1) and lung cancer (225 deaths; SMR, 1.1; 95% CI, 1.0–1.3) were slightly elevated. The risks for all cancers combined and for soft-tissue sarcomas and lymphomas increased with time since first exposure. In a direct comparison of persons exposed to higher chlorinated PCDDs and those exposed to lower chlorinated PCDDs or none, the rate ratio for all cancers combined was 1.3 (95% CI, 1.0–1.8). Increased risks were found for breast cancer in both women and men, endometrial cancer and testicular cancer (Table 6). The increased mortality rate from breast cancer was confined to female workers in the Boehringer plant in Germany (nine deaths; SMR, 2.8; 95% CI, 1.3–5.4). Two of three deaths from endometrial cancer similarly occurred among women who had worked in this plant. An excess risk was also seen for cancer of ‘other endocrine organs’, both deaths being from tumours of the suprarenal glands.

Table 6. Standardized mortality ratios (SMRs) for selected tumours in the 21 863 workers in the IARC international cohort study who had been exposed to phenoxy acetic acid herbicides or chlorophenols, by exposure to TCDD or higher chlorinated dioxins, 1939–92

Cause of death (ICD-9)

Exposed workers

Unexposed workers

 

No. of deaths

SMR

95% CI

No. of deaths

SMR

95% CI

All causes

2728

1.0

0.97–1.0

1367

0.91

0.86–0.96

All malignant neoplasms

710

1.1

1.0–1.2

398

0.96

0.87–1.1

Breast, female (174)

9

2.2

0.99–4.1

3

0.53

0.11–1.6

Breast, male (175)

2

2.6

0.31–9.3

0

0

0.00–7.7

Endometrium and uterus (179, 181, 182)

3

3.4

0.70–10

1

1.2

0.03–6.5

Ovary (183)

0

0

0.00–2.6

1

0.45

0.01–2.5

Prostate (185)

43

1.1

0.81–1.5

25

1.1

0.71–1.6

Testis (186)

4

1.3

0.36–3.4

3

1.3

0.28–3.9

Thyroid (193)

2

1.4

0.16–4.9

2

2.2

0.26–7.8

Other endocrine organs (194)

2

2.2

0.27–8.1

3

6.4

1.3–19

From Kogevinas et al. (1997); ICD-9, International Classification of Diseases, Ninth Revision; CI, confidence interval

Two nested case–control studies of soft-tissue sarcoma (11 incident cases, 55 controls) and non-Hodgkin lymphoma (32 incident cases, 158 controls) were conducted by Kogevinas et al. (1995) within the IARC cohort. A panel of three industrial hygienists estimated the exposure of the workers to 21 chemicals or mixtures, and a cumulative exposure score was calculated for each person and chemical, on the basis of estimated intensity and duration of exposure (in years). An excess risk for soft-tissue sarcoma was associated with exposure to any phenoxy acetic acid herbicide (odds ratio, 10; 95% CI, 1.2–91). Soft-tissue sarcoma was also associated with exposure to TCDD (odds ratio, 5.2; 95% CI, 0.9–32). The odds ratio for non-Hodgkin lymphoma and exposure to TCDD was 1.9 (95% CI, 0.7– 5.1). A monotonic increase in risk was observed with cumulative exposure to TCDD (Figure 13).

FIGURE 13

From Kogevinas et al. (1995)

Figure 13. Risk for non-Hodgkin lymphoma and soft-tissue sarcoma, with exposure to TCDD (from an exposure matrix), nested case-control study within the IARC cohort study

(c) Exposure of the general population and commercial pesticide and herbicide applicators

This review does not cover studies of phenoxy acetic acid herbicide applicators, including US Air Force personnel who applied Agent Orange, numerous community based case–control studies on soft-tissue sarcoma, malignant lymphoma and other neoplasms, and studies of persons exposed to unspecified combinations of pesticides and herbicides or to herbicides not contaminated by PCDDs (e.g. Axelson et al., 1980; Hardell et al., 1981; Blair et al., 1983; Coggon et al., 1986; Michalek et al., 1990; Coggon et al., 1991; Lynge, 1993). The available information indicates that, in these studies, the concentrations of TCDD were lower and, in most cases, considerably lower than that to which industrial workers and the population of Seveso were exposed. In one Swedish case–control study (Nygren et al., 1986), the mean serum concentration was 2 pg/g in 13 exposed persons and 3 pg/g in 18 unexposed persons. In a study of the most heavily exposed commercial sprayers in New Zealand (Smith et al., 1992), the concentration of TCDD was near the background level in serum of sprayers with 5–10 years of experience. A higher mean concentration of TCDD (53 pg/g) was found for sprayers who had first been employed before 1960 and with a mean duration of spraying 2,4,5-trichlorphenoxyacetic acid of 16 years. Lower values were observed among commercial sprayers in Australia (Johnson et al., 1992) and in US Air Force personnel who had sprayed phenoxyacetic acid herbicides in Viet Nam (Ketchum et al., 1999), the latter being near background.

The findings of studies of cancer risk among persons evaluated in community-based studies and of applicators are contradictory (Figure 14). The large discrepancies are probably due to misclassification of exposure, as most of the persons classified as exposed in these studies probably had TCDD burdens very similar to or only slightly higher than those of persons classified as unexposed.

FIGURE 14

From Johnson (1990)

Figure 14. Relative risks in 15 case-control studies of soft-tissue sarcoma and exposure to phenoxyacetic acid herbicides, chlorophenols and dioxins

One case–control study on liver cancer was conducted in Hanoi, Viet Nam. An increased risk for hepatocellular carcinoma (OR, 8.8; 95% CI, 1.9–4.1) was observed among persons who had been in military service in southern Viet Nam for 10 years or more during the time of spraying of Agent Orange by the US Air Force (Cordier et al., 1993).

(d) Cancer in children

The incidence of cancer among persons aged 0–19 years in Seveso, Italy, was analysed separately (Pesatori et al., 1993). Given the small number of cases, the three contaminated zones and the two sexes were grouped. Seventeen cases were observed (relative risk, 1.2; 95% CI, 0.7–2.1). Two cases of ovarian cancer were found, with none expected. Two thyroid gland cancers among girls gave a relative risk of 4.6 (95% CI, 0.6–33). The incidence of lymphatic and haematopoietic neoplasms was increased (nine cases; relative risk, 1.6; 95% CI, 0.7–3.4), and particularly those of Hodgkin lymphoma (three cases; RR, 2.0; 95% CI, 0.5–7.6) and myeloid leukaemia (three cases; relative risk, 2.7; 95% CI, 0.7–11).

(e) Conclusions

Low excess risks of the order of 50% were found for all neoplasms combined, in all studies of industrial cohorts in which exposure had been assessed adequately (Table 7). These excess risks were highly statistically significant, and any effect of chance can be excluded. The risk tended to be higher for workers with the heaviest exposure. Increased risks with time since first exposure were observed in those studies in which latency was evaluated (Kogevinas et al., 1997; Steenland et al., 1999). The risks for certain cancers were increased in some studies (lymphomas, multiple myeloma, soft-tissue sarcoma, lung cancer, liver cancer, breast cancer, testicular cancer, endometrial cancer), but, overall, the results were not consistent among studies, and no particular cancer appears to predominate.

Table 7. Mortality rates from all neoplasms in selected industrial cohorts with heavy exposure to polychlorinated dibenzodioxins and furans

Cohort (reference)

No. of deaths

SMR (95% CI)

IARC International cohort, 20 years since first exposure (Kogevinas et al., 1997)

394

1.2 (1.1–1.3)

Industrial populations: groups with heavy exposure

 

 

Workers with highest septile of exposure (Steenland et al., 1999)a

40

1.6 (1.2–1.8)

Cohorts I and II (Becher et al., 1996)

105

1.3 (1.0–1.5)

Cohort in plant with accident (Hooiveld et al., 1997)

51

1.5 (1.1–1.9)

Workers with chloracne 20 years after BASF accident (Ott & Zober, 1996)

18

1.9 (1.1–3.0)

SMR, standardized mortality ratio; CI, confidence interval

a SMR for whole cohort, 1.1 (95% CI, 1.0–1.2; 377 deaths)

In examining the findings on cancer risk in the most informative epidemiological studies, a number of issues should be noted. The excess risks observed were highly statistically significant, and any effect of chance could be excluded. In addition, the studies were prospective and were well conducted. Nevertheless, the results must be evaluated with caution, given that the overall risks are not very high and that the strongest evidence is for industrial populations. Furthermore, there are very few precedents of carcinogens that affect the risk for all cancers with no clear excess for any specific cancer (IARC, 1997). Finally, the strongest evidence comes from studies of persons whose exposure was two to three orders of magnitude higher than that of the general population. In order to extrapolate to the general population, models must be used, with the assumption of similar effects at high and low doses. At present, the real dilemma is not whether dioxins are or are not carcinogenic but, rather, the size of the risk associated with the very low exposure of the general population.

2.3.2 Effects other than cancer

Exposure to dioxins has been associated with a variety of adverse effects (Table 8), including chloracne and related skin lesions, meibomian gland dysfunction, peripheral neuropathy, developmental deficits and coughing and other symptoms of respiratory irritation. Most of the responses observed are biologically plausible, in the sense that experimental evidence also indicated such effects (Birnbaum & Tuomisto, 2000). Most of the available epidemiological studies, however, focused on mortality from cancer and were not designed to evaluate morbidity, e.g. neuropsychological effects, or transient effects such as changes in reproductive hormones. Epidemiological evidence exists for only few of these effects and, in contrast to the experimental evidence, is inconsistent for most effects other than cancer. The evidence in humans is currently conclusive only for dermatological effects and temporary increases in liver enzyme concentrations, and there is increasing evidence for an association with cardiovascular disease.

Table 8. Strength of epidemiological evidence for effects of TCDD other than cancer

Effect

Epidemiological evidence

Dermatological effects (chloracne)

Proven association

Gastrointestinal effects and liver enzymes

Temporary increases in liver enzymes proven

Cardiovascular disease and changes in lipid concentrations

Positive associations in most studies of heavy exposure, but results not entirely consistent: dose–response relationships in some studies

Diabetes

Overall, results not consistent; increased risks in population of Seveso and US Army herbicide sprayers in Viet Nam

Reproductive hormones and reproductive outcome

lnconsistent results for reproductive hormones; altered sex ratio of infants of heavily exposed couples in Seveso; no data on effects in women, e.g. endometriosis, fertility

Thyroid function

Some (small), inconsistent differences reported in T4, TSH, T4-binding globulin and T3 uptake

Neurological and psychological effects

Inconsistent effects reported in US Army herbicide sprayers and Seveso population (polyneuropathy, abnormal coordination); no association with depression

Respiratory system

Inconsistent evidence; irritation, reduced FEV1 and FVC in some studies.

Urinary system

No major renal or bladder dysfunction observed

Immunological effects

Inconsistent

T4, thyroxine; TSH, thyroid-stimulating hormone; T3, tri-iodothyronine; FEV1, forced expiratory volume in 1 s; FVC, forced vital capacity

(a) Reproductive effects

While there was clear evidence that dioxins and coplanar chemicals can cause a variety of adverse reproductive effects in experimental animals, the evidence for human populations was limited, except for the two episodes of rice oil poisoning in South-East Asia. Accidental consumption of rice oil contaminated with heat-degraded PCBs occurred in two separate episodes, in Japan in 1968 and in Taiwan (Province of China) in 1979. These incidents are referred to as ‘Yusho’ and ‘Yu-cheng’, which mean ‘oil poisoning’ in Japanese and Chinese, respectively. Large numbers of persons were exposed. In the Yusho incident, the average intake of the rice oil contaminated with PCBs, PCDFs and polychlorinated quaterphenyls was estimated to have been 150 ng/kg bw per day (as toxic equivalents), which is five orders of magnitude higher than the reported average background intake in several countries. Analysis of the rice oils in question showed that other chlorinated compounds, including polychlorinated terphenyls, naphthalenes, quarterphenyls and furans, were present, in addition to the PCBs Chen et al., 1985; Masuda et al., 1985).

A number of attempts have been made to determine whether US Air Force personnel involved in spraying TCDD-contaminated herbicides during the Viet Nam War had higher rates of adverse reproductive outcomes (for a review, see Environmental Protection Agency, 2000a). While there have been no strong indications of decreased fertility rates or an increased risk for adverse birth outcomes, such as spontaneous abortion, stillbirths or intrauterine growth retardation, some results suggest higher rates of birth defects, both total and specific (Centers for Disease Control, 1988; Michalek et al., 1998a). A number of the studies lack consistency and may have involved reporting bias or exposure misclassification (Wolfe et al., 1995).

When all birth outcomes (total, 2900) of women from the affected areas in Seveso in 1977–82 were reviewed, no significant increase in the frequency of total or specific birth defects was observed (Mastroiacovo et al., 1988). The authors noted that most of the participants were from outside the two zones with the heaviest TCDD contamination.

Both US Air Force personnel involved in spraying in Viet Nam and workers in the USA employed in factories producing TCDD-contaminated herbicides (chloro-phenoxyacetic acid derivatives) have undergone assessment for circulating gonadotropins and sperm characteristics. In the former group, current serum concentrations of testosterone, follicle-stimulating hormone and luteinizing hormone and testicular abnormalities, sperm count, sperm abnormalities and testicular volume were not associated with current or back-extrapolated serum TCDD concentrations (Henriksen et al., 1996). In 1987, the median serum TCDD concentration in this group was 13 pg/g of lipid (up to 620 pg/g) (Roegner et al., 1991). However, in the more heavily exposed group of herbicide production workers, linear regression analysis indicated that serum TCDD concentrations > 20 pg/g of lipid (upper quartile, 244–3400 pg/g) were positively associated with high serum concentrations of luteinizing and follicle-stimulating hormones and negatively associated with a low serum concentration of testosterone (Egeland et al., 1994).

In addition to the variety of developmental abnormalities observed in children born to mothers who consumed poisoned rice oil (Yusho and Yu-Cheng; see below), about twice as many women affected in the Yu-Cheng incident reported abnormal menstrual flow 14 years after poisoning when compared with neighbourhood control women (Yu et al., 2000). More Yu-Cheng-affected women also reported having had a stillbirth after 1979 (4.2%, with 1.7% in controls), although the overall fertility rates appeared to be aimilar.

Since the initial report that long-term exposure to low doses of TCDD was associated with endometriosis in non-human primates (Rier et al., 1993), attempts have been made to investigate the situation in humans. It has been pointed out (Bois & Eskenazi, 1994) that the level of exposure to dioxin of women in Seveso was comparable to the dose that caused endometriosis in rhesus monkeys. A retrospective study is under way to examine these effects in women in Seveso. Among 79 women attending an infertility clinic in Jerusalem, detectable concentrations of TCDD were found in the blood of eight of 44 with endometriosis (0.6–1.2 pg/g) and only one of 35 with mechanical infertility (0.4 pg/g) (Mayani et al., 1997). In Belgium, endometriosis was diagnosed by laparoscopy in 42 of 101 women who were undergoing treatment for infertility in 1995–98. Coplanar compounds were found in the blood (toxic equivalents, > 100 pg/g of serum lipid) of seven of these 42 women and only two patients who were infertile for mechanical reasons (Pauwels et al., 1999). After that initial report, an adjusted odds ratio for high serum toxic equivalents (presumably > 100 pg/g of lipid) and risk for endometriosis of 4.6 was reported (i.e. 4.6 times as many women with endometriosis would have high serum toxic equivalent values as determined by the CALUX bioassay in vitro) (Pauwels et al., 2000).

Further evidence for a link between TCDD and endometriosis is provided by experimental studies. Groups of eight ovariectomized nude mice injected intraperitoneally with human proliferative phase endometrial tissue which had been treated in vitro with 10–8 mol/l of estradiol and 10–6 mol/l of TCDD for 24 h developed more than twice as many lesions as tissue treated with 10–8 mol/l of estradiol alone (42 and 20, respectively). Furthermore, treatment of human endometrial tissue with TCDD completely blocked the antiproliferative effect seen when the tissue was pretreated with 5 x 10–7 mol/l of progesterone and estradiol (Bruner-Tran et al., 1999).

An initial report from Seveso indicated that the sex ratio of children born to parents who had lived in zone A, with the highest level of TCDD contamination, was altered in favour of females (Mocarelli et al., 1996). Of the births that occurred from 9 months after the accident to 7 years later, about twice as many children were female (26 males, 48 females). In nine highly exposed families (maternal and paternal serum TCDD concentrations > 100 pg/g of lipid), no male children were born (0 males, 9 females). During 1985–94, the sex ratio apparently reverted to normal (60 males, 64 females). In a follow-up study (Mocarelli et al., 2000), evidence was reported that a paternal serum concentration of TCDD > 80 pg/g of lipid and a body burden of 16–20 ng/kg bw was a significant predictor of a female birth. The most extreme modification of the sex ratio was observed among fathers who had been exposed when they were < 19 years of age (sex ratio, 0.38; 95% CI, 0.30–0.47). A biological explanation of this phenomenon has not yet been provided. General decreases in the normal male:female sex ratio of 1.06:1.0 have been observed in a number of industrialized countries since 1950 but not to the extent temporarily observed in Seveso (Davis et al., 1998).

Comparable analyses in other populations exposed to dioxins, PCBs or PCDFs (in studies of oil poisoning and herbicide spraying in Viet Nam) have not provided support for these findings (Michalek et al., 1998b; Rogan et al., 1999). However, the exposure of these populations to TCDD was considerably lower than that in Seveso. Of 137 births between 1978–85 to women who were identified from the Yu-Cheng registry, approximately equal numbers of males and females (68 and 69) were born (Rogan et al., 1999). Although women who had estimated body burdens of toxic equivalents of 2–3 ΅g/kg (up to two orders of magnitude higher than that of the background population) delivered children with a variety of developmental and neurological abnormalities, the heavy maternal (and presumably paternal) exposure to PCBs and coplanar contaminants appears to have had a limited effect on the sex ratio of their children.

(b) Developmental effects

Neurodevelopmental effects in children have been examined mainly in relation to exposure to mixtures of organochlorinated compounds, predominantly PCBs (Ribas et al 2001).

(i) Yusho and Yu-Cheng rice oil poisoning episodes

The most characteristic symptoms observed at birth in the infants of mothers certified as Yusho patients was a gray–dark-brown skin and gingival pigmentation, which usually resolved within 2–5 months. Other clinical features included natal teeth, abnormal skull calcification, ‘rocker-bottom’ heels and intrauterine growth retardation (Yamashita & Hayashi, 1985). Although limited follow-up has been reported of infants exposed in this incident perinatally, one group of 13 children has been described as being apathetic and hypotonic, with lower IQs, 7 years after the event (Harada, 1976).

In the larger Yu-Cheng episode of poisoning, children born up to 6 years after the outbreak (1979–85) weighed on average 0.5 kg less than age-matched neighbourhood controls. While the weight difference had resolved by the time the children were 6–13 years of age, they continued to be shorter (3.1 cm) than the control group (Guo et al., 1995). Behavioural and developmental assessment of this cohort has shown them to be hyperactive, with lower mean IQs and delayed cognitive and psychomotor development. Recent assessment of boys affected by the episode indicated they have short penises and alterations in spermatozoa, 12 youths aged 16–18 years having reduced sperm motility and increased abnormal sperm morphology when compared with 23 controls (Guo et al., 2000). A variety of adverse developmental deficits seen in the Yu-Cheng children were associated with maternal body burdens of ­ 2 ΅g/kg toxic equivalents, which were about two orders of magnitude higher than those of control populations.

The children exposed in the Yu-Cheng incident were assessed continuously from the age of 6 months to 15 years with one of four tools for measuring cognitive development (Lai et al., 2001). In comparison with a control group matched for age, residence and socio-economic status, Yu-Cheng children scored lower on a psychomotor development index at 2 years of age. At 4–5 years of age, the Yu-Cheng children had lower IQs (2–7-point difference on the Stanford-Binet test) than the controls, and they had lower scores on tests designed to evaluate general intelligence factors at 7, 8, 11 and 12 years of age. When the children reached 13 years of age, there was no significant difference in the test scores from those of the control group.

(ii) PCBs and dioxins in breast milk in The Netherlands

In a study of 418 mother–infant pairs enrolled in 1990–92, 209 infants were to be breastfed for at least 6 weeks and 209 infants were to be fed exclusively with formula. The median total toxic equivalents in the breast milk was 63 ng/kg of lipid (range, 25–160 ng/kg). The physical and neurological developmental of the infants was assessed from 2 weeks after birth to the present (42 months completed). Previous findings had indicated that prenatal exposure to dioxins and PCBs was associated with lower birth weight and reduced infant growth up to the age of 3 months, subtle alterations in thyroid hormone concentrations and immunological parameters and poorer neurological condition at 18 months of age (Cuijpers et al., 1997; Patandin et al., 1998, 1999; Boersma et al., 2000). The childrens’ intellectual functioning was assessed with the Dutch version of the Kaufman Assessment Battery for Children up to 42 months of age. Although higher PCB concentrations in maternal plasma (> 3.0 ΅g/l) were associated with lower cognitive scores in the entire cohort and in the formula-fed group, no effects on performance were seen of prenatal or lactational exposure (Patandin et al., 1999). All the scores for cognitive ability were within the range of those of the normal population. The immunological status of the children was also assessed at 42 months of age by administering a health questionnaire, measuring plasma antibodies and conducting lymphocyte marker analyses (Weisglas-Kuperus et al., 2000). The contributions of mono-ortho (14 ng/kg of lipid) and planar PCBs (15 ng/kg of lipid) toxic equivalents in breast milk were associated with a 10–17% higher incidence of recurrent (> six episodes) middle-ear infections, while the toxic equivalent of dioxin (36 ng/kg of lipid) was associated with a slightly (6%) higher prevalence of coughing, chest congestion or phlegm lasting for 10 or more days. There was no association between antibody titres, leukocyte counts and immunological markers and dioxin and/or PCB toxic equivalents.

(iii) Dental effects

In a study designed to investigate the causes of enamel defects in teeth, children born in Finland in 1987 underwent dental examinations at the age of 6–7 years and were classified according to the extent of changes in mineralization in their teeth. Although the duration of breastfeeding and breast milk toxic equivalents at 4 weeks of age (mean, 49 pg/g of lipid) were not associated with the changes in teeth mineralization, exposure to dioxin or furan toxic equivalents was associated with these changes (Alaluusua et al., 1996, 1999). In a related study, 202 children in a PCB contaminated area of Slovenia underwent a dental examination to assess developmental defects of the enamel (Jan & Vrbi, 2000). The tooth dentine of these children had more than five times more PCB than that of children in a control group (38 and 7 ng/g, respectively). On the basis of an analysis of maternal serum and milk samples and representative foods from the same region, the children were estimated to have been exposed to toxic equivalents at a rate of 39 pg/kg bw per day. About twice as many exposed children as controls were found to have permanent teeth with enamel defects on the labial surface (22% and 13%, respectively).

(c) Chloracne

Chloracne is the best recognized dermal effect of exposure to TCDD. Chloracne was reported in at least a few of the workers involved in the reported accidents at trichlorophenol production facilities. Chloracne has also been reported among workers involved in daily production of TCDD-contaminated products (Suskind & Hertzberg, 1984), such as phenoxyacetic acid herbicide workers, and in three laboratory technicians exposed to pure TCDD. Chloracne was diagnosed in at least 193 (0.6%) Seveso residents, mostly children, exposed during the accident (Table 9; Bertazzi et al., 1998). It is not clear whether the children were more susceptible or whether they had heavier exposure to TCDD. Of the populations with known exposure to TCDD, US Air Force personnel in Viet Nam and persons in Missouri, USA, who were accidentally exposed were not reported to have chloracne, but the latter were examined 10 years after exposure.

Table 9. Cases of chloracne diagnosed in children aged 3–14 years in an 8-month period after the Seveso accident

Contamination zone

Total population

Chloracne cases

Prevalence (%)

A

214

42

20

Part of zone A closest to factory

54

26

48

B

1 468

8

0.5

R

8 680

63

0.7

Non-A, -B or -R

48 263

51

0.1

Adapted from Bertazzi et al. (1998a)

Chloracne persisted in a few workers in plants in Germany and the USA many years after exposure. In Seveso residents, the condition disappeared after discontinuation of exposure, despite very high initial serum concentrations of TCDD, ranging from 820 to 56 000 pg/g, measured within 1 year of the accident.

Although chloracne is associated with the intensity of exposure to TCDD, there is no direct correlation between exposure and the occurrence of chloracne. Persons with severe chloracne are often found, on a group level, to have had heavier average exposure than those with moderate or no chloracne. This can clearly be seen in the analysis of TCCD concentrations of BASF workers exposed during the accident (Zober et al., 1997). However, an absence of chloracne is not equivalent to an absence of exposure, and the presence of chloracne is not necessarily equivalent to very heavy exposure.

Various other dermal effects other than chloracne have been described among persons exposed to TCDD, including red, irritated eyes, conjunctivitis and blepharitis, eyelid cysts, hyperpigmentation and hirsutism. These findings have been reported less frequently than chloracne.

(d) Gastrointestinal effects and effects on hepatic enzymes

Increased liver size was reported in workers in two trichlorophenol production plants, one in the former Czechoslovakia and the other in the USA. Temporary hepatomegaly was reported in five of 22 Seveso residents with severe chloracne (Reggiani, 1980), but hepatomegaly has not been observed in other populations of workers or in US Air Force personnel in Viet Nam.

lncreased activity of gamma-glutamyl transferase was observed in children in Seveso shortly after the accident, but this declined during the 5 subsequent years (Mocarelli et al., 1986). Persistently high gamma-glutamyl transferase activity has been observed in trichlorophenol production workers in various plants, and significantly high activity was observed in the US Air Force personnel (Roegner et al., 1991).

Increased serum activity of alanine and aspartate aminotrasferases is probably a transient effect of exposure. No increases were found in studies of persons 10–30 years after exposure. Clinical evidence of liver disease was not found in any of the studies in which increases were identified.

Increased concentrations of D-glucaric acid were found in adults and children in Seveso in 1976 (Ideo et al., 1985). By 1981, however, the concentrations were within the normal range. No increases were found in US Air Force personnel in Viet Nam (Roegner et al., 1991).

The metabolism of porphyrins was examined in two studies. In one study of trichlorophenol production workers (Bleiberg et al., 1964), porphyria cutanea tarda was reported in 11 of 29 persons with chloracne. In a study of industrial cohorts in the USA, no association was found between exposure to TCDD and the prevalence of porphyria cutanea tarda (Calvert et al., 1994).

(e) Effects on thyroid function

Thyroid function was associated with exposure to dioxin in some cohorts of production workers and in US Air Force personnel in Viet Nam, but most of the results were not statistically significant or consistent among studies (Sweeney & Mocarelli, 2000). In a study in a trichlorophenol production factory, no significant difference in T4 and T4-binding globulin concentrations was found between exposed and unexposed workers (Suskind & Hertzberg, 1984). In workers exposed in the BASF accident, the concentrations of TSH, T4 and T4-binding globulin were within normal levels; the concentrations of the latter two were positively correlated with that of TCDD (Ott et al., 1994). In the industrial cohort in the USA, no significant differences were found between exposed and unexposed persons, although the free T4 index and T4 concentration were increased in trichlorophenol production workers (Calvert et al., 1999). US Air Force personnel spraying herbicides in Viet Nam showed a slight, nonsignificant increase in TSH concentrations which was associated with TCDD concentration (Grubbs et al., 1995).

(f) Diabetes

Higher mean glucose concentrations were found among TCDD-exposed than among unexposed persons in the industrial cohort in the USA (Calvert et al., 1999), in US Air Force personnel in Viet Nam (Henriksen et al., 1997) and in workers exposed at the BASF facility at the time of the study, but not when the concentration was estimated at the time of last exposure (Ott et al., 1994). No association was found in workers in Nitro, West Virginia, USA (Suskind & Hertzberg, 1984). Among US Air Force personnel exposed in Viet Nam (Henriksen et al., 1997), the prevalence of diabetes and use of oral medication to control diabetes increased with increasing exposure to TCDD, while the time to onset of diabetes decreased with the intensity of exposure to dioxin (Figure 15). Among trichlorophenol production workers in the USA, the prevalence of diabetes was not associated with serum concentrations of TCDD. Men with very high TCDD concentrations (> 1500 pg/g), however, tended to have a high prevalence of diabetes (Calvert et al., 1999). In the industrial cohort in the USA (Steenland et al., 1999), any mention of diabetes on the death certificate was negatively associated with exposure. An increased rate of mortality from diabetes was seen among women who had been in zone A or B in Seveso (rate ratio, 24; 95% CI, 1.2–4.6), but no excess was observed among men (Pesatori et al., 1998).

FIGURE 15

From Henrikson et al. (1997)

Figure 15. Serum concentration of dioxin and prevalence of indicators of diabetes among veterans of the US Air Force defoliation operation in Viet Nam

(g) Cardiovascular effects and lipid concentrations

In the BASF accident cohort and the industrial cohort in the USA, no relationship was found between the concentration of total cholesterol or high-density and low-density lipoproteins and increasing serum concentrations of TCDD (Ott et al., 1994; Calvert et al., 1996). Similar results were found for Seveso residents (Mocarelli et al., 1986; Assennato et al., 1989). In contrast, a positive relationship between serum TCDD and total cholesterol concentration was found among US Air Force personnel in Viet Nam (Roegner et al., 1991). This association was less strong in a later follow-up study (Grubbs et al., 1995).

No or very small differences in concentrations of triglycerides were found in TCDD-exposed workers in the BASF accident cohort and the industrial cohort in the USA when compared with unexposed persons (Ott et al., 1994; Calvert et al., 1996), with similar results in Seveso (Mocarelli et al., 1986; Assennato et al., 1989). TCDD concentrations were consistently associated with triglyceride concentrations in US Air Force veterans of Viet Nam (Roegner et al., 1991; Grubbs et al., 1995).

Exposure to dioxins has been associated with an excess risk for heart disease. Excess mortality from ischaemic heart disease was found in several industrial cohorts (Table 10) and in zone A at Seveso. The study of non-flying US Air Force personnel gave mainly negative results, but there was an excess risk for personnel with the heaviest estimated exposure to TCDD. No excess risk was observed in an analysis of cardiovascular morbidity in a subset of the US industrial cohort. An increased risk with increasing exposure was observed in the Dutch production cohort (Figure 11) and the Boehringer cohort (Figure 10). In the US industrial cohort, the rate of mortality from heart disease showed a weak increasing trend with heavier exposure.

Table 10. Mortality rates from ischaemic heart disease in subgroups of occupational cohorts, US Air Force personnel in Viet Nam and the population of Seveso with heavy exposure to dioxins

Cohort (reference)

SMR (95% CI)

BASF accident (Ott & Zober, 1996)

0.6 (0.2–1.3)

IARC international cohort (Vena et al. (1998)

1.7 (1.2–2.3)

Dutch cohort (Hooiveld et al., 1998)

1.9 (0.9–3.6)

Boehringer cohort (Flesch-Janys et al., 1995)

1.4 (0.7–2.8)

US Air Force non-flying personnel (Michalek et al., 1998c)

1.5 (1.0–2.2)

US industrial cohort (Steenland et al., 1999)

1.8 (1.1–2.9)

Seveso, 15-year follow-up (Pesatori et al., 1996)

1.6 (1.2–2.5)

SMR, standardized mortality ratio; CI, confidence interval

(h) Neuropsychological effects

Numerous case reports have been made of an association between short- or long-term exposure to TCDD and headache, insomnia, nervousness, irritability, depression, anxiety, loss of libido and encepalopathy. There have been some reports of persistent symptoms. Effects were described in US Air Force personnel in Viet Nam (abnormal coordination), workers at the plant in Seveso (polyneuropathy in lower limbs) and in Seveso residents (neuropathy) (Pocchiari et al., 1979; Filippini et al., 1981; Roegner et al., 1991). No association was found between exposure to TCDD and depression in the US industrial cohort or in the US Air Force personnel (Roegner et al., 1991; Alderfer et al., 1992).

(i) Conclusions

The results of epidemiological studies on most health outcomes other than cancer are still inconsistent. The evidence is currently conclusive only for dermatological effects and temporary increases in liver enzyme activity, and there is increasing evidence for as association with cardiovascular disease, particularly among the most heavily exposed workers. Experimental data indicate that endocrine and reproductive effects should be among the most sensitive effects in both animals and humans. Only a few of these effects have been evaluated in epidemiological studies.

3. ANALYTICAL METHODS

3.1 Chemistry

PCDDs and PCDFs are tricyclic aromatic compounds with similar physical and chemical properties, and the two classes are structurally similar. In simplified terms, congeners of both classes are called ‘dioxins’, although they are differentiated chemically as dioxins and furans. The PCDDs comprise 75 individual compounds and the PCDFs comprise 135 compounds. Figure 16 shows the chemical structures of these classes. Of the 210 theoretically possible congeners, a subset of 17 has chlorine substitutions at the 2, 3, 7 and 8 positions. The most toxicologically potent dioxin is TCDD (Figure 17). Table 1 lists all 17 congeners with chlorine substitution in the 2,3,7,8 position, and all these congeners are included in this assessment.

FIGURE 16

Figure 16. Chemical structures of polychlorinated dibenzo-para-dioxins (PCDDs) and polychlorinated dibenzofurans (PCDFs)

FIGURE 17

Figure 17. Chemical structure of 2,3,7,8-tetrachlorodibenzodioxin (TCDD)

The most important properties of PCDDs and PCDFs are their chemical stability and high solubility in fat and organic solvents, with low solubility in water. Because of these properties, PCDDs and PCDFs are primarily associated with particulate and organic matter in environmental samples, and they accumulate in the food chain.

PCBs have been manufactured as technical products, which are liquids with variable viscosity. These liquids are mixtures of dozens of different compounds (‘congeners’) depending on the degree of chlorination.

A total of 209 congeners are theoretically possible. Following IUPAC rules, Ballschmiter & Zell (1980) developed systematic numbering of all congeners, which is now used worldwide for identification of a specific congener. Four octachloro-biphenyls were numbered incorrectly and corrected subsequently (Schulte & Malisch, 1983; Deutsche Forschungsgemeinschaft, 1988).

The physical and chemical properties of each congener vary according to the degree and position of chlorine substitution. Figure 18 shows the chemical structure of PCBs.

FIGURE 18

Figure 18. Chemical structure of polychlorinated biphenyls (PCBs)

From a chemical and toxicological point of view, PCBs can be divided into three groups: with no chlorine substitution in the ortho position (non-ortho PCBs, coplanar PCBs), with one chlorine in the ortho position (mono-ortho PCBs) and di-ortho-substituted congeners. A WHO consultation (van Leeuwen & Younes, 2000) included 12 of the 209 PCBs in the set of coplanar compounds (see Table 1). The PCBs considered to have coplanar properties are those with either one or no chlorine substitution in the ortho position. The toxicity of these congeners is different from that of other PCBs. Figure 19 shows the planar structure of 3,3΄,4,4΄-TCB as an example of a non-ortho PCB, which is coplanar, as both rings are in the same plane. Figure 20 shows the twisted structure of di-ortho-substituted PCBs.

FIGURE 19

Figure 19. Structure of a non-ortho (coplanar) PCB

FIGURE 20

Figure 20. Structure of a di-ortho-substituted PCB

The mono-ortho PCB 2,4,4΄-trichlorobiphenyl and the di-ortho PCBs 2,2΄,5,5΄-tetrachlorobiphenyl, 2,2΄,4,5,5΄-pentachlorobiphenyl, 2,2΄, 3,4,4΄,5΄-hexachloro-biphenyl, 2,2΄,4,4΄,5,5΄-hexachlorobiphenyl and 2,2΄,3,4,4΄,5,5΄-heptachlorbiphenyl (and also sometimes the mono-ortho PCB 2,3΄,4,4΄,5-PeCB) are called ‘marker PCBs’ and are used as decisive congeners in determining tolerances for PCBs in various matrices in some countries. They represent the subsets of low- and high-chlorinated PCBs, as they are indicative congeners that can be determined readily. However, coplanar PCB content cannot be derived by determination of these marker PCBs. A variety of PCB products was available with an unknown range of contamination with coplanar PCBs. Except in the case of accidents, food samples are contaminated with PCBs from different sources as a result of various transport and bioaccumulation processes. For these reasons, information on the occurrence of PCBs in general (e.g. estimatates of total PCBs or of marker PCBs) is not sufficient to describe exposure to the coplanar PCBs. Similarly, it is not known whether regulations for total PCBs adequately address coplanar PCBs.

As with PCDDs and PCDFs, the most important properties of PCBs are their chemical stability and high lipophilicity. Because of these properties, PCBs are also primarily associated with particulate and organic matter in environmental samples and accumulate in the food chain.

3.2 Description of methods

3.2.1 Congeners to be determined

PCDDs and PCDFs are usually found as complex mixtures of variable composition in various matrices. Their identification and quantification require highly sophisticated analysis, as the 17 toxic congeners with 2,3,7,8-chlorine substitution must be separated from the less toxic congeners. Usually, PCDDs and PCDFs are determined by capillary gas chromatography (GC) with mass spectrometry (MS).

In the past, analyses of PCBs focused mainly on determination of total PCBs or the marker congeners 2,4,4΄-trichlorobiphenyl, 2,2΄,5,5΄-TCB, 2,2΄,4,5,5΄-PeCB, 2,2΄,3,4,4΄,5΄-HxCB, 2,2΄,4,4΄,5,5΄-HeCB and 2,2΄,3,4,4΄,5,5΄-HpCB, which are the predominant PCB congeners in humans and foods of animal origin. However, these PCB congeners appear to have relatively little toxicity. On the basis of the available toxicological information, the non-ortho PCBs 3,3΄,4,4΄-TCB, 3,4,4΄,5-TCB, 3,3΄,4,4΄,5-PeCB and 3,3΄,4,4΄,5,5΄-HxCB and the mono-ortho congeners 2,3,3΄,4,4΄-PeCB, 2,3,4,4΄,5-PeCB, 2,3΄,4,4΄,5-PeCB, 2΄,3,4,4΄,5-PeCB, 2,3,3΄,4,4΄,5-HxCB, 2,3,3΄,4,4΄,5΄-HxCB, 2,3΄,4,4΄,5,5΄-HxCB and 2,3,3΄,4,4΄,5,5΄-HpCB were assigned TEFs by a WHO expert group in 1997 and are therefore analysed to determine their PCB toxic equivalent. Data on these coplanar PCB congeners are still scarce. As non-ortho PCBs occur in a much lower range of concentrations, mono-ortho and non-ortho PCBs must be determined separately. Reliable determinations of non-ortho PCBs in food have been performed by high-resolution MS, as demonstrated in interlaboratory studies.

3.2.2 Availability of official methods

In contrast to methods for the determination of food additives, residues (pesticides, veterinary drugs) and many contaminants, there are no standardized or harmonized methods (‘official methods’) for the determination of dioxins or coplanar PCBs in food. Reliable results can be obtained in the absence of official methods if the method used is shown to fit the purpose and to fulfill analytical quality criteria developed in other fields of residue analysis.

3.2.3 Upper- and lower-bound concentrations

For comparison of analytical results with regulatory limits and, in general, with results from other laboratories, the limit of detection (LOD; lowest limit for qualitative identification) and/or the limit of quantification (LOQ) must be taken into account. For analysis of PCDDs and PCDFs, all 17 congeners with 2,3,7,8-substitution must be determined. To calculate toxic equivalents, the results for each congener are multiplied by the specific TEF and then added up. In most cases, the concentrations of a few of the 17 congeners are below the LOD and/or the LOQ. The situation can become critical if many congeners cannot be determined or if the toxicologically important congeners are not found.

Various approaches are used to address this problem:

The LOD or LOQ can become a critical factor in decisions based on analytical results if many congeners are not determined or if congeners representing the higher TEFs are not found. This is because the value of ‘undetected’ congeners is needed in order to estimate the overall toxic equivalents. For example, TCDD and 1,2,3,7,8-PCDD have a TEF of 1, while other, more prevalent congeners have TEFs of 0.1 or 0.01. If analysis of a sample results in detection of large quantities of TCDD or 1,2,3,7,8-PCDD, the resulting toxic equivalents will be affected by the method used to estimate the values. Some laboratories calculate the contribution of undetected congeners to toxic equivalents as zero, whereas others use the full LOD or the full LOQ to estimate toxic equivalents. Thus, estimates of dioxin content may have low or high bias. The effect of these biases is obscured by the reporting of a single value for toxic equivalents.

An example of the effect of using replacement values for undetected congeners in samples is shown in Table 11. It is clear that the method used to estimate a value for undetected congeners can dramatically affect the reported toxic equivalents value for a food. If 0 is used, low estimates of dioxin content may result, owing either to truly low concentrations in the sample or to high LODs/LOQs that did not allow quantification of the congeners with higher TEFs. For example, Table 11 shows that the imputation method has little effect on the estimated toxic equivalents for beef Stroganoff when high-resolution MS is used. However, with ion trap MS, with which the LOD is 5–10 times higher, use of 0 to equate to nondetection results in an artificially low estimate of toxic equivalents when compared with that achieved with high-resolution MS.

Table 11. Variation in estimated toxic equivalents according to method of analysis and method of imputation, selected food samples collected in the USA in 1999 (values in pg of toxic equivalents per gram in whole food)

Food

Ion trap mass spectrometry

High-resolution mass spectrometry

 

ND equal to

Ratio high/low estimate

ND equal to

Ratio high/low estimate

 

0

LOD/2

LOD

LOQ

0

LOD/2

LOD

LOQ

Cream cheese

0.15

0.23

0.30

0.59

4

0.081

0.099

0.12

0.19

2

Cheddar cheese

0.17

0.25

0.33

0.65

4

0.14

0.15

0.16

0.21

2

Frankfurters

0.086

0.22

0.34

0.34

4

0.14

0.18

0.21

0.86

6

Corn chips

0.0013

0.15

0.29

0.87

669

0.0012

0.037

0.07

0.22

183

Beef Stroganoff, homemade

0.00006

0.058

0.12

0.35

5833

0.023

0.031

0.039

0.072

3

English muffin, plain

0.029

0.12

0.20

0.55

19

0.045

0.080

0.11

0.25

6

ND, not detected; LOD, limit of determination; LOQ, limit of quantification

Table 11 illustrates the importance of using analytical methods with low LODs when making decisions about tolerances. Even with the more sensitive high-resolution MS, however, the estimate of toxic equivalents for corn chips in Table 11 was strongly influenced by the choice of replacement value for undetected compounds, with a 183-fold difference when zero and the LOQ are used. Thus, it is important not only to be able to use high-resolution MS for determination but also to use an appropriate amount of sample for extraction and clean-up to obtain a sufficient concentration in the final sample for use in GC–MS. The explanations, conclusions and recommendations in sections 3.2.4 and 3.2.5 below indicate the analytical requirements to avoid using methods with insufficient sensitivity.

To ensure that a finding of low concentrations of dioxin is really the result of low concentrations in the sample, the concept of tolerances as upper-bound concentrations has been developed in some countries and for some uses of estimates of concentrations. This concept requires use of the full LOD or LOQ instead of zero for undetectable substances: Upper-bound concentrations are calculated on the basis of the assumption that all the values for various congeners that are below the LOD/LOQ are equal to the LOD/LOQ.

When the LOQs are high relative to the decision criteria for congeners, therefore adding significantly to the estimated toxic equivalents, use of the upper-bound LOQ can result in artificially high toxic equivalents. This should be considered in defining background contamination, monitoring tolerances or estimating intake. In methods with insufficient sensitivity, the difference between lower-bound and upper-bound concentrations may be 10–100 or, in extreme cases, even higher. For example, if the sensitivity of a method is inappropriate for monitoring a tolerance, use of the concept of upper-bound LOD leads to estimates of toxic equivalents that are ‘false-positive’ results. This is a clear indication that a more sensitive method is needed. In particular, use of low-resolution MS in analysing food or samples of low weight or quantity (for a quick, easy analysis) can result in relatively high values for dioxin content as the upper-bound LOD. This bias cannot be seen in reported toxic equivalents, unless results for individual congeners are available. Thus, in defining background contamination or evaluating exposure, published data must be reviewed critically to eliminate relatively high values that are the result simply of inadequate LODs.

For setting and monitoring tolerances on the basis of toxic equivalents, the closeness of the LOQ to the appropriate tolerance must be evaluated as part of a decision to accept or reject a food. High LODs relative to the tolerance should lead to rejection of an analytical result for a sample on the basis of poor quality assurance, and consequently poor reliability of the estimate of toxic equivalents. Therefore, some governments may choose to apply upper-bound estimates of toxic equivalents, with a preference for the upper-bound LOQ rather than the upper-bound LOD, as a screening method in order to remove questionable foods from the marketplace. In the absence of these steps, there is a risk that foods in which a maximum level of a toxicant is exceeded will reach consumers. It is the responsibility of laboratories to achieve the required sensitivity in order to avoid unnecessary rejection of analytical results or of foods.

In risk assessment, use of upper-bound concentrations leads to overestimes of intake and use of lower-bound concentrations to underestimates. For this purpose, imputation of half the LOD for undetected congeners yields an acceptable estimate of both toxic equivalents and its associated standard deviation of uncertainty (Hoogerbrugge & Liem, 2000).

The Committee therefore recommended that, in future, laboratories report their results in relation to the lower-bound, upper-bound and half the LOD. In that way, all the necessary information is available for interpretation of the results according to specific requirements. As a minimum, it must be clear from a report which concept was applied.

3.2.4 Required sensitivity and analytical approaches

Whereas many environmental samples (such as soil and sewage sludge) can be analysed by low-resolution MS, feed, food and human milk or tissue samples should be analysed for ‘ultra-trace’ concentrations (usually, 0.1–1 pg/g of lipid in milk and meat and in eggs of caged chickens, as toxic equivalents; mean, 10 pg/g of lipid in wild and farmed freshwater fish or > 100 pg/g of lipid in cases of (highly) elevated concentrations, as toxic equivalents; 0.1–0.5 pg/g dry matter for food of vegetable origin, as WHO toxic equivalents). As the fat content of foods of animal origin varies widely, a wide range of values for PCDD/PCDF is calculated on a fresh weight basis. Therefore, lipid-adjusted toxic equivalents are used for dioxins in animal foods, in order to provide a uniform basis for setting tolerances. For fish, dioxin content should be reported on the basis of both the fat content and on fresh weight, in view of the extremely wide range of fat contents of various kinds of fish.

For reliable analysis of food samples with contamination in the range of that of the normal background, high-resolution MS has been shown to have the required sensitivity and specificity. In collaborative studies for the determination of PCDDs/PCDFs in various foods, use of high-resolution MS was successful (Chemisches Landes- und Staatliches Veterinäruntersuchungsamt Münster, 1996; Malisch et al., 1996, 1997, 2000b; Lindström et al., 2000). This was also the basis for selecting reference laboratories for WHO-coordinated studies of exposure to dioxins in human milk (see section 6.4.2).

The required specificity can also be provided by tandem MS (MS/MS). While MS/MS with sector or quadrupole instruments requires a series of mass analysers in space, ion traps require one mass analyser to perform MS/MS in time. As the techniques for MS/MS depend on the type of mass analyser used, a variety of instruments has been designed. The advantage of ion-trapping MS/MS systems is their low price. Nevertheless, their sensitivity is considerably lower than that of high-resolution MS instruments. The LOQ for ion-trapping MS/MS systems for TCDD (signal:noise, 3:1) can be assumed to be in the range 100–300 fg, whereas that of modern high-resolution MS instruments is about 3 fg. The reduced sensitivity can be compensated to a certain degree by using much larger samples for extraction and clean-up. However, use of 10-fold (or more) larger samples causes problems with regard to the availability of sample material and the analytical procedure. Therefore, ion-trapping MS/MS could be used as screening method for selecting high concentrations, like bioassays. However, unlike bioassays, this screening method makes it possible to see patterns of congeners.

Several reviews (Cooke et al., 2000; Hilscherova et al., 2000; Hoogenboom et al., 2000; Behnisch et al., 2001a,b) and guidelines (reviewed by Cooke et al., 2000; Behnisch et al., 2001a) have been published on new techniques for measuring dioxin toxic equivalents. Bioassays such as CALUX (chemical-activated luciferase gene expression), Ah immunoassays, EROD bioassays and enzyme immunoassays have been developed for rapid screening of various matrices (food, sediments, soil, fly ash). So far, only CALUX has been used for food. In contrast to MS methods, these methods allow biological interpretations. While GC–MS is the most powerful method for identifying and quantifying congeners and for congener-specific pattern recognition, it does not provide a direct measure of total dioxin-like toxicity (toxic equivalents) for all congeners in a matrix that act through the Ah receptor pathway. The TEF concept is used to transform the GC–MS database into results relevant to toxicity, i.e. the individual concentrations of those congeners that have been assigned a TEF is multiplied by the TEF, and these are added to give the total toxic equivalents. The bioassays themselves provide an indication of the total toxic equivalents of dioxin-like activity present in a certain matrix, including possible interactive (synergistic or antagonistic) effects of all the coplanar congeners in a complex mixture. However, the bioassays in their present form cannot discriminate between different classes and/or congeners of PCBs, PCDDs and PCDFs, partly depending on the clean-up procedure, and therefore cannot reveal the pattern of congeners.

A ‘toxicity identification evaluation’ approach (Environmental Protection Agency, 1992) has been proposed, to give a ‘real world picture’ of which compounds are most responsible for the dioxin-like activity of the sample, on the basis of differences in polarity in fractionation and selection of stable and unstable compounds. This approach may allow detection of novel coplanar compounds auch as bromodioxins or combinations of bromo–chlor dioxins. It is a means for identifying patterns in different compound classes rather than single congeners, but it requires further developments and is not considered useful for screening for PCDD/Fs and coplanar PCBs.

Hoogenboom et al. (2000, 2001) reported that the CALUX has been validated in their laboratory for milk fat and citrus pulp pellets to monitor adherence to limits of 5 pg/g of lipid and 500 pg/kg of product, respectively, as toxic equivalents. The results for other animal and plant fats were comparable to those for milk fat, and those for feed ingredients (with the exception of kaolinitic clay) were similar to the results for citrus pulp. The CALUX method is sensitive not only to dioxins but also to other Ah-receptor agonists, possibly resulting in ‘false-positive’ results. This potential problem was partly overcome by use of selective clean-up procedures (such as acid silica) and long exposure of the cells (24 h). However, as the possibility of ‘false-negative’ results appears to be negligible, the method was considered ideal for screening samples with no dioxins and samples suspected of containing dioxins and requiring further investigation by the GC–MS reference method.

The results of a comparison of simultaneous analysis of 19 samples (eight fats, three feeds, four eggs, four milks) by CALUX and GC–MS were presented by van Overmeire et al. (2000). A strong correlation was found from a double-log graph in the range 0.1–10 000 pg/g of lipid as toxic equivalents, although the double-log presentation reveals some differences in the numerical values. According to the authors, the observed correlation indicates that analyses by CALUX are predictive of the results of GC–MS, and the method can be used as a rapid, sensitive screen for the dioxin content of feed and food samples. Concentrations of 5 pg/g of lipid, as toxic equivalents, were concordant with the two methods. A tolerance of 5 pg/g lipid, as WHO toxic equivalents (only for PCDD/PCDF), was set by Belgium during the recent incident of dioxin contamination for milk, eggs and meat samples, in order to exclude highly contaminated samples from being marketed.

The usual background concentration of contamination of food of animal origin (except fish from European waters) is 0.1–2 pg/g of lipid, as toxic equivalents; most samples contain < 1 pg/g of lipid, as WHO toxic equivalents (only for PCDD/PCDF). If maximum concentrations are set, they will be based on concentrations above the average background. Therefore, a method should be developed to allow screening for elevated concentrations (e.g. above a fixed maximum or action level) or as a routine or reference method to determine the content of PCDD/PCDF and/or coplanar PCBs in the range of the usual background contamination or target values. This is true for all kinds of methods of detection, whether GC–MS or bioassays.

3.2.5 Conclusions and recommendations

These examples show that the principles common to all analytical methods are valid for dioxins as well. First, the purpose of a method should be defined clearly, including the matrix to be analysed, the content to be determined reliably, possible limitations and whether it is to be used for quick screening or reliable determination (routine or reference method). Then, it should be shown that the method is suitable for the purpose, by validation and demonstration of basic minimal statistical requirements. Last but not least, the applicability of the method should be proven in collaborative studies.

As mentioned above, there are no official methods for the determination of PCDDs/PCDFs or coplanar PCBs in food. Each laboratory has its own method, although it must demonstrate that the method is suitable for the purpose. Nevertheless, the criteria for accepting methods for determining dioxin in food must be harmonized. Given the wide variety of methods and the unequal quality assurance and control across laboratories, harmonization is needed to allow free trade if maximum levels are developed. These principles should be valid for both GC–MS methods and for bioassays. Joint papers by 13 authors in nine different international institutions contain a discussion of these general considerations and of GC–MS methods (Malisch et al., 2001), and papers by 11 authors in 10 institutions address the bioassays (Behnisch et al., 2001c).

(a) Maximum limits, action levels, target levels

Legislative measures for dioxins in food can comprise ‘maximum limits’, ‘action levels’ and ‘target levels’. Maximum limits can be set at a strict but feasible concentration in order to allow elimination of products contaminated at an unaccep-tably high level. Action levels can be set to allow monitoring of increased concentrations. Target levels could be set at a concentration that would result in an ultimate dietary intake below the lower range recommended by WHO. Some governments have begun to discuss such levels, the ratio of maximum limits to action levels being about 1.5 and that of maximum levels to target levels being about 5.

If legislative measures are based on these three concepts, analysis for certification must allow relaible determination of the dioxin content in the range of maximum levels and action levels. For evaluation of exposure and time trends, analysis must be oriented to the target levels.

(b) General requirements for monitoring adherence to maximum levels

General statistical parameters have been established for the analysis of other residues which might provide orientation. For example, to allow certification (in national or international commerce) relative to maximum levels, laboratories should be able to meet certain basic requirements, such as:

Studies from laboratories in which high-resolution GC with high-resolution MS are used have shown that all 17 PCDD/PCDF congeners with 2,3,7,8-substitution can be determined reliably, even at concentrations < 1 pg/g of lipid, as WHO toxic equivalents (only PCDD/PCDF included). However, successful participation in inter-calibration studies for e.g. soil or sewage samples does not necessarily prove competence in the analysis of food samples, with their lower range of contamination. Therefore, continuous participation in interlaboratory studies for determination of dioxins and coplanar PCBs in the relevant food matrices is mandatory.

(c) General requirements for monitoring background contamination

As long as no target values are fixed, the following requirements should be met for reliable determination of concentrations in the range of the usual background contamination:

(d) Special requirements for GC–MS methods

(e) Harmonized quality criteria for cell-based and kit-based bioassays

(i) General

Before biological or chemical analyses are begun, the relevant quality control criteria should be well defined. The characteristics of these criteria will vary, depending on the analytical approach being used. Three analytical approaches can be used in bioassays:

(1) The first is a screening approach, in which the response of samples is compared with that of a reference sample at the action limit. Samples with a response less than that of the reference are considered to be negative, while those with a response higher than that of the reference are considered suspect. The requirements may be less strict than those for a quantitative method:

(2) The second is a quantitative approach, which requires a standard dilution series, duplicate or triplicate clean-up and measurement, and blank and recovery controls. The result should be expressed as toxic equivalents (pg/g), thereby assuming that the compounds responsible for the signal correspond to the TEF principle. This can be done by using TCDD (or a standard mixture of dioxins and furans) to produce a calibration curve for calculating the WHO toxic equivalents in the extract and thus in the sample. This is subsequently corrected for the WHO toxic equivalents calculated for a blank sample (which may include impurities from the solvents and chemicals used) and recovery (calculated from the WHO toxic equivalents in a reference sample around the limit of the residue). It should be noted that part of the apparent loss to recovery may be due to differences between TEF values in the bioassays (relative potency) and the official TEF values set by WHO.

(3) The third design involves use of bioassays for various toxicological end-points, e.g. Ah receptor or antibodies. In this case, a wide range of concentrations should be used to evaluate a full median effective concentration (EC50) and to obtain toxic equivalent values from various measures (EC50, EC10 or lowest data point closest to the minimal LOQ).

(ii) General quality criteria for bioassays

(iii) Special requirements for all bioanalytical detection methods

(iii) Special requirements for cell-based bioassays

(iv) Special requirements for kit-based bioassays

The standard quality criteria requirements for kit-based bioassays could include the following (see e.g. EPA method 4025 or 4035):

Essentially, kits consist of screening assays, and many of the same quality criteria guidelines suggested for screening assays should be followed.

(f) Combination of bioassays and GC–MS analysis

The next level of screening should involve more detailed characterization. A positive sample (with toxic equivalents) should be characterized by chemical analysis, involving fractionation of the sample and identification of the responsible coplanar compounds and then analysed in a bioassay. The correspondence should be reported, as discussed above.

Samples for which the results of bioassays for toxic equivalents cannot be explained by chemical characterization should be studied further in vitro or in vivo by other techniques.

3.3 Sampling protocols

There are no specific guidelines for sampling protocols for food samples to be analysed for their dioxin content. Therefore, basic rules for sampling for organic contaminants or pesticides should be followed. The primary requirement is a representative, homogeneous laboratory sample with no secondary contamination.

3.3.1 Personnel

A qualified, authorized person should perform sampling.

3.3.2 Representative sample

Samples must be representative of the lots or sublots from which they are taken. Compliance with maximum levels or action levels should be established on the basis of the concentrations determined in the laboratory sample.

Lots are identifiable quantities of food delivered at one time and determined by the official to have common characteristics, such as origin, variety, type of packaging, packer, consigner or markings. In the case of fish, they should be of comparable size. Sublots are designated parts of a large lot to which the sampling method is applied. Each sublot must be physically separate and identifiable. An incremental sample is a quantity of material taken from a single place in a lot or sublot. As far as possible, incremental samples should be taken at various places distributed throughout the lot or sublot. An aggregate sample is the combined total of all the incremental samples taken from the lot or sublot. It should be at least 1 kg, unless impractical. A laboratory sample for the purposes of enforcement, trade and refereeing should be taken from the homogenized aggregate sample, unless this conflicts with States’ regulations on sampling. The sample used to ensure enforcement should be large enough to allow at least duplicate analysis.

3.3.3 Packaging, transport and storage of aggregate and laboratory samples

Each aggregate and laboratory sample should be placed in a clean, inert container offering adequate protection from contamination, loss of analytes by adsorption to the internal wall of the container or damage in transit. Glassware offers good protection from contamination and can be cleaned easily. Polyethylene and polypropylene containers also provide protection against damage during transit. Containers made from halogenated substances (such as polyvinylchloride) are not considered suitable for this purpose. Although dioxins are chemically stable, samples must be stored and transported in such a way that the food sample does not deteriorate. In particular, the fat content should not be changed, for example by microbiological or enzymatic processes, as the content of the compounds in food of animal origin is generally calculated on a fat basis.

3.3.4 Human milk samples

WHO has recommended addition of K2Cr2O7 tablets to human milk samples during collection of portions and for transport in the third round of studies of exposure, if freezing of the portions cannot be guaranteed. This helps to avoid microbiological deterioration of the samples. If the portions can be frozen immediately after collection and the collected portions can be shipped in a frozen state, addition of K2Cr2O7 tablets is unnecessary.

3.3.5 Sealing and labelling

Each sample taken for official use should be sealed at the place of sampling and identified following States’ regulations. A record must be kept of each sampling, permitting each lot to be identified unambiguously and giving the date and place of sampling, together with any additional information likely to be of assistance to the analyst.

3.3.6 Edible parts

For determination of dioxins in food, only the edible parts are analysed. Vegetables should be washed with water to separate them from adhering soil.

4. EFFECTS OF PROCESSING

As dioxins are chemically stable, lipophilic substances, no changes in dioxin content with processing would be expected. However, studies have been performed to identify possible changes during processing or preparation of food.

4.1 Carry-over from packaging and filter material

At the end of the 1980s, transfer of PCDDs/PCDFs from paper into milk was found in milk samples packaged in cardboard containers. Large amounts of 2,3,7,8-TCDF and the presence of 1,2,7,7-TCDF and other less important TCDFs gave a characteristic pattern of contamination. The migration of these substances from paper through polyethylene foil into milk was clearly demonstrated. Additionally, carry-over of PCDDs/PCDFs and alkylated chlorodibenzofurans from coffee filter paper, from bleaching of pulp and paper with chlorine, into brewed coffee has been detected. Measures were taken to change the process, and significant reductions in the dioxin content were achieved.

4.2 Smoking, frying and broiling

Smoking of meat or fish samples can increase the dioxin and furan content depending on the smoking conditions (Körner & Hagenmaier, 1991; Mayer, 1998; Mayer & Jahr, 1998).

Information on the levels of PCDDs, PCDFs and coplanar PCBs in food relate primarily to uncooked food, although possible changes in PCDD/PCDF congener content and toxic equivalents after cooking have been reported. Broiling of hamburger samples resulted in an approximately 50% decrease in the total toxic equivalents (wet weight) per hamburger, but the decrease appeared to be due solely to the decrease in wet weight associated with loss of water and loss of PCDD/sPCDFs with the fat (Schecter et al., 1996).

In further studies, it was shown that the total toxic equivalents (PCDDs/PCDFs and PCBs) in hamburger, bacon and catfish decreased by an average of 50% as a result of broiling. However, the concentration remained the same in hamburger, increased by 84% in bacon and decreased by 34% in catfish (Schecter et al., 1997). On average, the total measured concentration (pg/kg whole weight) increased by 14% in hamburger and by 29% in bacon and decreased by 33% in catfish (Schecter et al., 1998b).

In a study of the effect of pan-frying beef patties, the PCDD/PCDF concentration was reduced by 40–50% by cooking. Most of the reduction was accounted for by the amount in fat liberated from the patties during cooking. There was nevertheless an overall deficit of 6–14% for each congener (Petroske et al., 1997, 1998). The authors attributed the losses to volatility, degradation and processing, but errors due to analytical imprecision or loss of fat were not considered. The general physical characteristics of these compounds would suggest that volatility and degradation are unlikely reasons for loss during cooking or broiling. Nevertheless, pan frying of ground beef significantly reduces the amounts of PCDD/PCDFs as consumed, provided that the fats and juices are not eaten.

After beef patties were cooked by several methods, each PCDD/PCDF congenr could be completely accounted for on a mass balance basis (within the experimental variability of the method). There was no indication of either loss or formation of PCDDs/PCDFs during the cooking process (Thorpe et al., 1999).

No de-novo synthesis of dioxins was observed after deep frying of scallops of pork covered with egg and crumbs either with or without salt and pepper. The frying temperature was high (180 °C). These results were confirmed by analysis of fats used for deep frying from a large hotel. Only a balance between the dioxin content of the raw scallops and the fat was observed (Schwind & Hecht, 2000).

It is therefore unlikely that PCDDs/PCDFs are formed or lost during usual cooking processes. Changes in dioxin content can be seen on a fresh weight basis owing to changes in fat and water content. As a whole, the dioxin content on a mass balance basis is expected to be constant in meat, fat and juices; however, changes can occur between these phases.

5. LEVELS AND PATTERNS OF CONTAMINATION
OF FOOD COMMODITIES

5.1 Results of surveys

Data were submitted by six countries (Belgium, Canada, Japan, New Zealand, Poland and the USA) and by the Commission of the European Union (2000a) (see Table 12).

Table 12. Summary of data submitted to JECFA on dioxins and PCBs, with number of results (and number of individual samples in parentheses)

Country

Milk and products

Meat and products

Fish and products

Eggs

Vegetable products

Fats and oils

PCDDs/PCDFs

 

 

 

 

 

 

Belgium

49 (49)

0

0

0

0

0

Canada

65 (65)

100 (100)

25 (25)

5 (5)

0

10 (10)

European Union

65 (2543)

56 (606)

80 (6281)

13 (1300)

21 (110)

1 (8)

Japan

36 (36)

110 (110)

190 (187)

9 (9)

84 (84)

0

New Zealand

4 (75)

9 (188)

3 (75)

1 (15)

2 (75)

1 (20)

Poland

5 (48)

3 (95)

9 (45)

1 (12)

0

2 (12)

USA

79a (357)

11 (170)

188 (323)

22 (486)

0

0

PCBs

 

 

 

 

 

 

Belgium

20 (20)

0

0

0

0

0

Canada

65 (65)

100 (100)

25 (25)

5 (5)

0

10 (10)

European Union

18 (434)

11 (114)

24 (536)

3 (156)

16 (23)

1 (8)

Japan

36 (36)

110 (110)

190 (186)

9 (9)

84 (84)

0

New Zealand

4 (75)

9 (188)

3 (75)

1 (15)

2 (75)

1 (20)

USA

38a (38)

10 (223)

2 (2)

2 (4)

1 (NR)

0

NR, not reported

a Individual results from the USA for milk represent a sample from one collection station or a composite from 51 collection stations.

In all countries in which substantial numbers of samples have been analysed over time, the concentrations of dioxin in food were decreasing up to the end of the 1990s. In several countries, this decrease was slowed, or even partly reversed in some food categories, due to contamination of feed. In addition, at the end of the 1990s, the measurements initiated at the beginning of the decade did not show the same strong reduction. For the international intake assessment reported here, only data collected after 1995 were considered. These represent 13 589 food products pooled in 1256 samples for PCDDs/PCDFs and 2636 food products pooled in 796 samples for PCBs.

There were large differences in the amount, detail and quality of the data from the participating countries. In particular, the results appeared to have been obtained largely without adequate harmonization of analytical procedures and/or intercalib-ration among the laboratories in different countries. This ultimately affects the comparability of the results. In addition, some of the data may be overestimates of the dioxin concentration, owing to lack of sufficient sensitivity for determination in the analytical laboratories where the assays were performed.

Most data were submitted for PCDDs/PCDFs and PCBs, as a sum of the congeners, weighted by the WHO TEFs. In some cases, an upper-bound approach was used with the LOQ in calculating toxic equivalents for those congeners that were not quantified. In other cases, the concentration of the contaminant may have been underestimated, when a lower-bound approach, with a zero value for those congeners that were not quantified, was used. As the Committee did not have access to the original analytical results, all the concentrations were expressed as the sum of congeners.

As it was impossible to identify analytical results for targeted samples, all the data were considered representative of total contamination of foods. Some data were not used because the mean level of contamination was not quantified.

Belgium: Belgium submitted information about the occurrence of dioxins in pooled milk and cheese samples, using the GEMS/Food electronic submission format. Data on 49 individual samples analysed in March, August and October 2000 were included.

Canada: Canada submitted the results of a total diet study conducted between 1992 and 1995 on dioxins and PCBs in 220 composite foods collected in five cities. All the results submitted were for individual samples. Data from older studies conducted between 1980 and 1989 were also submitted but were excluded from the international intake assessment.

Japan: Japan provided results on the occurrence of PCDDs/PCDFs and PCBs in fish and fish products, meat and meat products, various fruits and vegetables and milk. All the results provided were for individual samples on both a fresh weight and lipid basis, allowing estimation of the distribution of contamination.

New Zealand: New Zealand submitted the results of analyses for 53 foods purchased in 1997 in five cities. The foods were grouped into 22 composites and analysed to determine the concentrations of PCDDs/PCDFs and 23 PCB congeners including non-ortho PCBs.

Poland: Poland submitted the results of determinations of PCDD/PCDF content in Polish products. As there were not enough results to represent the eastern European region, they were not used in the international intake assessment.

USA: A draft report from the EPA was submitted, which contained a large amount of information collected in the USA and other countries. Additional information was provided by the Food and Drug Administration.

Commission of the European Union: A report of Scientific Cooperation (SCOOP) Task Force 3.2.5, entitled Assessment of Dietary Intake of Dioxins and Related PCBs by the Population of European Union, Member States, was submitted (Commission of the EuropeanUnion, 2000a). Ten countries (Belgium, Denmark, Finland, France, Germany, Italy, The Netherlands, Norway, Sweden and the United Kingdom) provided data on the occurrence of PCDDs/PCDFs and coplanar PCBs in food products and human milk. Samples were obtained nationally from rural and industrial sites in the period 1982–99.

The SCOOP database comprises information reviewed previously (IARC, 1997; AEA Technology (1999), but it also contains more recent material resulting from studies conducted up to the end of 1999. In addition, it contains information on coplanar PCBs that was not taken into consideration in the other reports.

5.2 Distribution in foods

For most countries, there were not enough individual data to allow generation of a full curve for the distribution of concentrations, as they were submitted in an aggregated format. As agreed at the FAO/WHO workshop on exposure assessment to contaminants (WHO, 2000), aggregated data were weighted as a function of the number of initial samples. Each result was multiplied by the number of individual samples in the original survey, and the sum of the products was then divided by the total number of individual samples to obtain a weighted mean of the contamination of foods by PCDDs/PCDFs and PCBs.

National data were therefore aggregated by region when sufficient data were available (western Europe, North America, Oceania, Far East; Table 13). The data were not sufficient to permit a realistic estimate of the distribution of contaminants for the rest of the world.

In a second step, a log-normal distribution of contaminants in foods was assumed, and the distribution was modelled from the weighted mean and a geometric standard deviation of 3.0 derived from the data on concentrations, for six broad groups of foods: meat and meat products, eggs, fish and fish products, milk and milk products, vegetable products and fats and oils. The geometric standard deviation was determined after log-transformation of the data by the following formula:

where N is the total number of individual samples, ni is the number of individual samples in a pooled result x, xi is the national pooled result, and x is the mean of the national pooled results.

The percentiles of these distribution curves were determined, and the median values (50th percentiles) are presented in Table 13.

Table 13. Weighted mean and derived median of concentrations of PCDDs, PCDFs and coplanar PCBs in six food groups, expressed as WHO toxic equivalents (pg/g whole food)

Region or country

Food category

PCDDs/PCDFs

Coplanar PCBs

 

 

Weighted mean

Derived median

Weighted mean

Derived median

Western Europe

Dairy

0.07

0.04

0.08

0.07

 

Eggs

0.16

0.15

0.07

0.06

 

Fish

0.47

0.31

2.55

0.90

 

Meat

0.08

0.06

0.41

0.08

 

Vegetable products

0.04

0.03

0.04

0.00

Japan

Dairy

0.06

0.04

0.04

0.02

 

Eggs

0.07

0.03

0.06

0.04

 

Fish

0.37

0.11

0.69

0.19

 

Meat

0.09

0.01

0.04

0.009

 

Vegetable products

0.003

0.002

0.02

0.003

New Zealand

Dairy

0.02

0.02

0.01

0.008

 

Fish

0.06

0.05

0.09

0.07

 

Meat

0.01

0.01

0.02

0.01

 

Vegetable products

0.008

0.008

0.00

0.00

North America

Dairy

0.10

0.07

0.02a

0.01a

 

Eggs

0.17

0.14

0.04a

0.02a

 

Fish

0.56

0.28

0.13a

0.08a

 

Meat

0.13

0.10

0.14a

0.05a

All

Fats and oils

0.21

0.10

0.07a

0.02a

a Data on PCBs frequently did not include mono-ortho PCBs

The data used to construct the distributions of coplanar PCBs are of much lesser quality than those for PCDDs/PCDFs. The distribution in meat in western Europe is based on data from only two countries, The Netherlands and Sweden. Additionally, the North American data and the data on fats and oils frequently did not include mono-ortho PCBs, resulting in an underestimate of the total concentration of WHO toxic equivalents of coplanar PCBs. In general, Table 13 shows that the median concentrations (derived from the modelled distributions) of coplanar PCBs in most food groups are of the same order of magnitude as those of the PCDDs/PCDFs. The region with the highest concentrations in most food groups is western Europe. The Committee recognized that there were significant differences within the food categories in Table 13 and that the data used in this analysis may not reflect the true mean for a food category. For example, the mean concentrations of PCDDs, PCDFs and coplanar PCBs and the consumption rates vary considerably for different fish species, and it is not possible to verify that the mean shown in Table 13 represents the fish species most commonly eaten. The data received were not sufficient to allow an analysis to account for this variation. Furthermore, the median concentrations of coplanar PCBs in fish were considerably (1.5–3 times) higher than those of PCDDs/PCDFs.

6. ESTIMATED DIETARY INTAKE

6.1 Introduction and background

Dietary intake of PCDDs, PCDFs and coplanar PCBs can be estimated in one of three ways. The most direct estimate is probably that made by measuring concentrations in duplicate diets collected over a certain time (e.g. Liem et al., 1997). Intake can also be estimated by analysing human tissues or by using a pharmaco-kinetics model in an inverse way, i.e. estimating dose from measured concentrations in blood, milk or adipose tissue (Pinsky & Lorber, 1998). Section 6.4 provides an overview of data available from various biomonitoring programmes, with special attention to breast milk.

Although these methods are promising, dietary intake of PCDDs, PCDFs and coplanar PCBs is usually estimated from data on food consumption and concentrations measured in foods and food groups. The latter method is also recommended by WHO (2000), and was used in the assessment described below. National monitoring programmes and surveys and food consumption surveys provide data on concentrations and food consumption, which were used here (section 6.2.3). Intake was also evaluated on the basis of the GEMS/Food regional diets (section 6.2.4).

Maximum limits for concentrations of contaminants in food groups are often proposed as regulatory instruments to exclude contaminated food products from the food chain. Section 6.3 illustrates the effects of maximum limits on the percentage of products that exceed these limits for a given distribution of concentrations in products. One of the potential benefits of such limits, a reduction in intake, is also discussed.

6.2 Calculations of intake

6.2.1 Relevant period of intake

PCDDs, PCDFs and coplanar PCBs (referred to as ‘dioxins’ below) have long half-lives in the body and therefore accumulate during continuous exposure. This property of dioxins has several implications for the period of intake of relevance to this assessment. First, the concentration of dioxin in a person’s blood (or the internal concentration of dioxin to which a target organ is exposed) will rise over time as more dioxin is ingested, reaching a pseudo-steady state only after decades. Second, after exposure stops, the decrease in the body’s stored dioxin (and the decrease in exposure of internal organs) will be similarly slow, only half of the accumulated dose disappearing within about 7 years (see section 2.1). Third, because of the long storage of dioxin in the body and the consequent daily exposure, a person’s daily ingestion of dioxin will typically have a very small or even negligible effect on their overall exposure. It is for these reasons that the appropriate averaging period for evaluating the intake of coplanar compounds is months or even years.

Furthermore, it is important to note in this regard that short-term variation in the concentration of dioxin in a food has much less effect on overall exposure and risk than it might for other food contaminants. For example, dioxin contamination that causes an even 100-fold higher concentration in a typical meal would result in a relatively small increase (< 3%) in the total body burden and a relatively small increase in risk with each meal consumed (see section 2.1). Similarly, when the concentration remains constant, the effect of changing food consumption with age will have a limited effect on long-term exposure. These considerations are illustrated in Figure 21, which shows the food consumption of the Dutch population in grams per capita per day as a function of age. The Figure indicates that the lower food consumption of young people contributes only modestly to the total food consumption during a lifetime (area under the curve). The long-term mean food consumption rate can therefore be approximated from the mean adult consumption rate.

FIGURE 21

Derived from the Dutch National Food Consumption Survey 1998 (Voedingscentrum, 1998) with regression analysis and nested analysis of variance (Slob, 1993). Percentiles indicate the between-person variation in each age group

Figure 21. Long-term mean food consumption per capita by the Dutch population as a function of age

These considerations lead to the conclusion that, for dioxins, the long-term mean intake is of relevance. This can be estimated from the long-term mean food consumption of the population and the mean concentrations of PCDDs, PCDFs and coplanar PCBs in foodstuffs. (See Appendix 1 for a mathematical synopsis.)

6.2.2 Methods used

The following definitions were adopted:

Dietary intake: The dietary intake of PCDDs, PCDFs and coplanar PCBs is defined as the amount of these contaminants that is ingested in food per unit time. Dietary intake is expressed in one of two ways. The EPA of the USA usually expresses intake in picograms of toxic equivalents per capita per unit time (day), while in most European countries intake is expressed in picograms of toxic equivalents per kilogram body weight per unit time (day or week). The latter measure requires that data on body weight be available. In the current assessment, intake is expressed as picograms of toxic equivalents per capita per day.

Region: When applied to concentrations and diets, an area comprised of individual nations or other geopolitical units that are likely to have separate food sources and markets but common dietary characteristics. The five GEMS/Food regional diets fit this definition.

Between-person variation: When applied to intake and food consumption, variation between individuals in a population within a nation or other geopolitical unit that is likely to have common food sources and markets.

Between-country variation: When applied to concentrations, variation between long-term mean concentrations in specific food groups in areas that predominantly do not share food sources or markets. National boundaries are assumed to define these populations within an acceptable degree of error for this analysis.

Within-food variation: When applied to concentrations, variation between consumed portions of a given food group during the period considered in the analysis. For example, the within-food variation in dioxin concentration for the group ‘Fish’ would comprise the variation in dioxin concentration from one meal to the next during the period of exposure (lifetime or other) of that individual. This variation is composed of variation due to differences between species and variation related to differences between fish of the same species. The within-food variation in dioxin concentration is assumed to be equivalent to the between-sample variation for the samples considered for each food group in this analysis.

(a) Intake calculations

Mean intake per person can be calculated from mean food consumption, the composition of the diet and the mean concentrations of PCDDs, PCDFs and coplanar PCBs in food from a local market, as follows:

where JT is the long-term mean personal intake of a contaminant (pg of WHO toxic equivalents per day), Ci is the mean concentration (in different portions) of the contaminant in food group i (pg of WHO toxic equivalents per g whole food), If is the mean food consumption (g/day), N is the number of food groups considered and fi is the fraction of food group i that contributes to total food consumption. (See also Appendix 1.)

The contribution of a food group i to the total intake of PCDDs/PCDFs or coplanar PCBs is obtained from the partial intake, Ji for group i, as follows:

If the mean food consumption per person and the mean concentration are considered random variables, it becomes possible to evaluate the distribution of the dietary intake by a certain population. The approach used corresponds to the method for assessing intake of contaminants and toxins in food recommended by a FAO/WHO workshop on the topic (WHO, 2000). In short, the procedure below was followed (see Appendix 1 for a mathematical synopsis):

  1. Concentration distributions indicating between-country variation in the mean were constructed for various regions and food groups from the data on occurrence submitted by the countries. These distributions were assumed to be log-normal (see section 5).
  2. Data on food consumption were used to estimate mean consumption and the between-person variation in food consumption in different diets. Log-normal food consumption distributions were constructed for each diet. Additionally, the contribution of the recognized food groups to total food consumption in these diets was derived from data on food consumption (see (d) and (e) below).
  3. The dietary intake of a particular population was assessed by combining the concentration and the food consumption distributions for that population by Monte Carlo simulations (with 10 000 trials for each simulation). In each Monte Carlo trial, dietary intake was estimated by multiplying random realizations of food consumption and concentrations in different food groups sampled from their distributions (Eq.1). The concentrations were weighted according to the fraction that each food group contributes to total food intake (Eqs 1 and 2). The collective intake estimates obtained by the Monte Carlo approach thus form a distribution of dietary intake for each population studied. The distributions are characterized by the median intake and two high percentiles (80th and 90th). Note that the 90th percentile is a realistic estimate corresponding to the mean intake that is exceeded by 10% of the population considered.

(b) Compounds

Dietary intake was calculated for the sum of dioxins (PCDDs and PCDFs) and for the sum of all coplanar PCBs, weighted according to their TEFs. The results are expressed as WHO toxic equivalents (van den Berg et al., 1998). Calculations could not be performed for the sum of all coplanar compounds, because data on the occurrence of PCDDs/PCDFs and coplanar PCBs were obtained independently.

(c) Distributions of concentrations

Data on occurrence were submitted by various countries (see section 5), and these data were used to compile regional distributions of concentrations, each distribution characterized by two parameters, the median (Table 13) and a geometric standard deviation (GSD). Several types of products were censored from the distributions in order to match them with the available data on food consumption.

The censored data included game, offal, liver and pure animal fats for meat; butter for dairy products; and hepatopancreas for shellfish. A second point is that intake for a region can be estimated only with full coverage of the distributions of concentrations for all product groups. Table 13 shows that this was not the case for North America or Oceania. In order to construct complete data series for each region, missing data are replaced by data for the closest region (western European data for North America, Far Eastern data for Oceania).

Essentially, the variation in concentrations within each food group and region consists of a ‘between-country’ and a ‘within-food’ component.

(i) Between-country variation

In calculating intakes, only the between-country component of the variation in concentration is relevant. Such variation implies that each region has areas with less and more contaminated areas. In other words, it is assumed that persons living in a country where there is higher contamination will not ‘dilute’ their daily intake of dioxin by eating food from a country where there is less contamination, nor will persons living in a less highly contaminated area frequently consume foods from a more contaminated area. The between-country variation was estimated from the results submitted for this assessment, which consisted mainly of means of aggregated data, i.e. measurements in pooled samples or means of series of individual measurements (see section 5). The GSD of a data series of means thus refers to the variation between those means.

It is nevertheless difficult to establish an accurate estimate of the GSD because the data consist of means, and the underlying sampling volume and the number of samples differ for each result. In the data on occurrence, the mean number of samples per result in the different regions ranged from 1 to 100 with GSD values of 1.2–10. The range of GSDs for the larger numbers of samples per result (> 20) was 1.2–3.4. On the basis of these results, a universal between-country GSD of 3 was assumed.

(ii) Within-food variation

The within-food component of the variation represents variation in concentrations in different portions of one food group bought in one area. This component is not used in intake estimates, because the long averaging time for dioxin intake renders ‘meal-to-meal’ variation irrelevant to the consideration of long-term risk. It is assumed that consumers choose food randomly with respect to the distribution of concentra-tions of contaminants and will therefore have an intake over time that is an approximation of the true mean of that distribution. (See also Appendix 1.)

An accurate estimate of the GSD for within-food variation requires measurement of concentrations in a set of individual products within one food group and one area. Since data on occurrence were not submitted for this purpose, the within-food variation was derived from Hoogerbrugge et al. (2000), who estimated the within-food variation among individual samples in five data sets for different food groups in The Netherlands. The relative standard deviations ranged from 0.38 to 1.2. On the basis of these results, a universal within-food GSD of 2 was assumed for this study. The within-food variation is not needed for calculating intake (see above); however, calculations of non-compliance require quantification of this variation (see (f) below).

(d) National diets

Four national diets were defined, on the basis of data on food consumption in France (Agence française de Sécurité Sanitaire des Aliments, 1999), the Netherlands (Voedingscentrum, 1998), the United Kingdom (Ministry of Agriculture, Fisheries and Food, 2000) and the USA (Department of Agriculture, 1997). For each diet, the mean consumption of the sum of the relevant food groups is specified. The average fraction that each group contributes to total consumption was derived from national food consumption surveys. Use of the average ignores interindividual variation in the composition of a diet; however, as broad food groups are evaluated, this generalization is appropriate. The between-person variation in total food consumption is assumed to be homogeneous throughout the world, giving a GSD of 1.3. This value was derived from the Dutch National Food Consumption Survey (Voedingscentrum, 1998; Figure 21) and extrapolated to the other diets. It must emphasized that the GSD for the food consumption curves accounts for interindividual differences in long-term consumption patterns.

(e) International diets

Calculations were also performed for the GEMS/Food regional diets (WHO, 1998). These diets are not derived from data on food consumption but from food production, import and export balances, as summarized by the FAO in their Food Balance Sheets. Comparison with the detailed results of national food consumption surveys shows that this type of data on food consumption provides estimates that are more than 15% higher than actual mean food consumption (WHO, 1998). The GEMS/Food regional diets were characterized similarly to the national diets.

(f) Hypothetical maximum limits

Concentrations of PCDDs, PCDFs and coplanar PCBs vary between individual products bought from retailers. This variation is indicated as the ‘within-food’ variation (see section (c) above). Maximum limits for concentrations of contaminants in food groups are often proposed as regulatory instruments to exclude (highly) contaminated food products from the food chain. For example, Belgian legislation specifies maximum limits for PCDDs and PCDFs in poultry, beef, pork, eggs and milk (Belgisch Staatsblad, 1999).

An estimate of the theoretical effect of such limit values on non-compliance and on reducing intake is required in order to evaluate their potential effectiveness. Non-compliance is defined as the fraction of products that does not comply with a maximum limit, i.e. the percentage of products containing concentrations above than that limit. The reduction in intake associated with this maximum limit can be estimated by recalculating the mean concentration and intake for a food group after the tail of the distribution curve for concentrations in individual products has been truncated. This was done by combining the analytical model presented in Appendix 2 with the probabilistic approach shown in Appendix 1.

A cost–benefit analysis can therefore be performed for alternative sets of maximum limits. The costs are represented by the percentage of products that exceeds the maximum limit, and the benefits are represented by the reduction in intake. A comprehensive derivation of the equations used to calculate non-compliance and the mean concentration of a truncated distribution is presented in Appendix 2.

6.2.3 National estimates of intake

(a) Central trend

The estimated intakes of PCDDs/PCDFs in the national diets are presented in Table 14 for three European countries and the USA. Studies of food consumption were available from these countries, from which national diets could be compiled that covered all six relevant product groups. The results in Table 14 indicate that the estimated median intakes from these diets differ only slightly (range, 66–84 pg per capita per day).

Table 14. Statistical descriptors (median, 80th (P80) and 90th (P90) percentiles) of estimated distributions of intake (pg of toxic equivalents per capita per day) in national diets

Source of data on concentrations

Diet

PCDDs/PCDFs

Dioxin-like PCBs

Median

P80

P90

Median

P80

P90

Western Europe

France

80

140

190

94

180

250

Western Europe

Netherlands

66

120

160

60

120

160

Western Europe

United Kingdom

78

140

180

81

150

220

North America

USA

84

150

210

17

34

50

A difference exists, however, between the estimates provided here and estimates made recently by national agencies of the intake of PCDDs/PCDFs. In the exposure assessments submitted by various countries, the following medians were reported: Germany, 32 pg per capita per day (Hecht & Blüthgen, 1998); The Netherlands, 55 pg per capita per day (Liem et al., 1996); United Kingdom, 48 pg per capita per day (Food Standards Agency, 2000) and the USA, 41 pg per capita per day (Environmental Protection Agency, 2000a). The SCOOP report on dietary intake (Commission of the European Union, 2000a) provides an overview of intake in Europe in both recent and older surveys. The more recent estimates of mean intake of PCDDs/PCDFs are in the range 29–97 pg per capita per day (inter-national toxic equivalents).The estimates provided in the current assessment (Table 14) are about 75% higher than these national estimates. This can be explained largely by methodological differences. The national estimates are based mainly on the results of monitoring, representing a wide range of possible concentrations in foods. The data on concentrations submitted for the current assessment, however, originated from a wide range of sources, including monitoring studies but also surveys. As surveys focus on products that are expected to contain high concentrations, use of these data may result in an overrepresentation of high values. A second explanation for the difference is that the national estimates often differed in their consideration of products of plant origin, which can contribute 6–45% of the total intake (Commission of the European Union, 2000a). Another factor may be the effect of undetectable levels and the way in which these are dealt with in the various intake assessments (see section 3.2.3). These points show that intake estimates must be interpreted with caution.

The estimated intakes of coplanar PCBs differ little from those of PCDDs/PCDFs (Table 14). National estimates of the intake of coplanar PCBs intake were available from several countries: Netherlands, 71 pg per capita per day (Liem et al., 1996); United Kingdom, 54 pg per capita per day (Food Standards Agency, 2000). As for PCDDs/PCDFs, the national intakes are lower than those calculated in the current exercise. The range of intakes of PCBs reported in the SCOOP report (Commission of the European Union, 2000a) was 57–110 pg per capita per day (PCB toxic equivalents), which corresponds better to those calculated here.

The estimated intakes from the European diets suggest that the contributions of dioxins and PCBs to the total intake of toxic equivalents are approximately equal. The estimated intake (17 pg per capita per day) in the diet in the USA is considerably lower than those in the European diets. This may be due partly to the absence in the North American concentration distribution of a contribution from mono-ortho PCBs to the toxic equivalents. The most recent national estimate for the USA was 24 pg per capita per day (Environmental Protection Agency, 2000a). Similar caution should be exercised in interpreting the estimates of intake of coplanar PCBs and for PCDDs/PCDFs.

(b) High intakes

High intakes within a population are considered to be those at the 80th and 90th percentiles (Table 14). These percentiles in the simulated intake distributions are the result of combining random realizations from the distributions of food consumption and concentration with simultaneously high values. They thus represent consumers who eat large amounts and live in areas where relatively more heavily contaminated products are sold.

The 90th percentile value is about 2.5 times higher than the median value. This difference corresponds roughly to the ratio of high to mean intakes reported in the SCOOP assessment (Commission of the European Union, 2000a).

(c) Dietary sources

Table 15 lists the mean contributions of various food groups to the intake of PCDDs/PCDFs. The contributions of oils and fats (3–11%) and eggs (4–5%) are relatively small, as would be expected from their proportion in relation to total food consumption. Vegetable products make a relatively large contribution (24–31%), which is larger than might be expected from the concentrations of dioxins in these products. It is due by the fact that 40–50% of food consumed consists of vegetables, cereals and fruit. The contributions of the other groups are: 21–35% from dairy products, 15–32% from meat and 6–19% from fish. These estimates correspond well with the SCOOP assessment (Commission of the European Union, 2000a). Table 15 also provides some insight into the effect of national eating traditions on the intake from each food group. For example, greater contributions were made from meat and fish products in France, from dairy products in The Netherlands and from fish and dairy products in the United Kingdom, although the same distributions of concentrations were used, but the diets are different.

Table 15. Mean contribution (%) of food groups to intake

Diet

Vegetable products

Oils and fats

Fish

Eggs

Dairy products

Meat

PCDDs/PCDFs

France

24

4

19

5

21

28

Netherlands

28

11

7

4

35

15

United Kingdom

31

6

16

4

26

16

USA

28

3

6

5

26

32

Coplanar PCBs

France

0

1

43

2

25

30

Netherlands

0

3

20

2

55

21

United Kingdom

0

1

42

1

35

20

USA

0

3

8

3

15

71

Table 15 also shows the mean contributions of the food groups to the intake of coplanar PCBs. As for the PCDDs/PCDFs, the contributions of oils and fats and eggs are relatively small. In contrast to the PCDDs/PCDFs, the contribution of vegetables is negligible. In the European countries, however, the contribution of fish to the intake of coplanar PCBs (20–42%) is higher than for PCDDs/PCDFs (6–19%). In the diet in the USA, the contribution of meat is striking. This results should, however, be interpreted with caution, since mono-ortho PCBs were not fully included in the North American concentration distribution.

6.2.4 Regional estimates of intake

(a) Central trend

The estimated median intakes from the GEMS/Foods regional diets are listed in Table 16. The concentration distributions used in the assessment are those for western Europe (SCOOP data), the Far East (data from Japan), North America (data from Canada and the USA) and Oceania (data from New Zealand). The estimated intakes of PCDDs/PCDFs and coplanar PCBs in the GEMS/Foods regional diets are considerably higher than those for the national diets: the results in Tables 13 and 16 indicate that the estimated median intake in the GEMS/Foods European diets are 45% and 60% higher than in the national diets. This is due mainly to the fact that food consumption is overestimated in the GEMS/Foods diets, as they are based on food production balances instead of food consumption.

Table 16. Statistical descriptors (median, 80th (P80) and 90th (P90) percentiles) of estimated distributions of intake (pg of toxic equivalents percapita per day) in GEMS/Foods regional diets

Source of data on concentrations

Diet

PCDDs/PCDFs

Dioxin-like PCBs

 

 

Median

P80

P90

Median

P80

P90

Western Europe

European

110

190

250

110

210

300

North America

European

140

230

310

27

50

70

Oceania

European

35

56

72

20

33

44

Far East

Far Eastern

13

22

30

13

25

37

Although the estimates of intake obtained from the GEMS/Foods diets are probably not accurate, they allow a comparison among regions (see Table 16). The lowest estimates are for the Far East and Oceania, which is due mainly to the lower concentrations than in Europe and North America in most food groups. However, the estimates for the Far East and Oceania are based on data from only two countries, Japan and New Zealand, respectively. The much lower result obtained when the data on concentration from Japan were used is probably not very accurate, as the difference is not consistent with the observed concentrations of coplanar compounds in breast milk samples from Japan (Environment Agency of Japan, 1999), which correspond to those in Europe and North America.

For the coplanar PCBs, a high median estimated intake was computed from the western European concentration distribution and the European regional diet. The estimated intakes in the other regions are at least 75% lower than that for western Europe. The results for North America must be interpreted with caution, as the underlying concentration distribution does not include the contribution of mono-ortho PCBs to toxic equivalents (see Table 13). When the result for North America is ignored, the median intakes of PCDDs/PCDFs and coplanar PCBs for the other regions are approximately equal.

(b) High intakes

High intakes within a population are considered to be those at the 80th and 90th percentiles (Table 16). On the basis of the model, the consumers represented by these percentiles eat large amounts of food and live in areas where products with relatively greater contamination are sold. The 90th percentile is about 2.5 times the median value, which results from the combined between-person variation in food intake and the between-country variation in the concentration distributions.

(c) Dietary sources

Table 17 shows the mean contributions of various food groups to the intake of PCDDs/PCDFs. As in the national diets, the contribution from eggs (4–7%) is relatively small, while the contribution from vegetable products (17–30%) is larger than would be expected from the concentrations of dioxins in these products. Table 17 illustrates some interesting differences between the regions: The results suggest that the contributions from oils and fats to the diets in the Far East (18%) and Oceania (22%) are larger than in other regions. This is because of the low concentrations of PCDDs/PCDFs in the remaining food groups in these regions when compared with the distributions in oils and fats, which was assumed to be of global scope. The contribution of fish to the intake of PCDDs/PCDFs in the Far East was much higher than in the other regions (41% and 9–19%) because of the higher concentrations in fish than in meat and dairy products in this region and the relatively greater amount of fish in the Far Eastern diet than in other regions .

Table 17. Mean contribution (%) of food groups to intake in the GEMS/Food regional diets

Diet

Vegetable products

Oils and fats

Fish

Eggs

Dairy products

Meat

PCDDs/PCDFs

 

 

 

 

 

 

Far East

17

18

41

4

13

6

Oceania

28

22

9

4

27

10

North America

24

5

14

5

27

23

Western Europe

30

7

19

7

20

17

Coplanar PCBs

 

 

 

 

 

 

Far East

18

4

64

5

5

4

Oceania

16

9

24

11

19

21

North America

0

6

19

3

17

55

Western Europe

0

1

49

3

27

20

Table 17 also lists the mean contributions of various food groups to the intake of PCBs. The contributions from eggs and oils and fats are relatively small. The high contribution from fish is clearly related to the relatively high concentrations of PCBs in this food group (Table 13). This has the greatest effect on the Far Eastern diet, which contains more fish than the other diets. The large contribution from meat in the North American diet is also striking, although this result should be interpreted with caution, as the concentration distribution of coplanar PCBs is incomplete (Table 13).

6.3 Effects of regulatory limits

6.3.1 Dioxin concentrations

The theoretical sensitivities of non-compliance and means in truncated distributions to different maximum limits can be visualized from application of the equations in Appendix 2. Figure 22 shows the effect of maximum limits of 1–8 pg/g of fat on non-compliance, given the original mean concentration. For example, in a certain country, the mean concentration of dioxins in eggs is 4 pg/g of fat (x-axis). For a maximum limit of 6 pg/g of fat (see the 6 pg/g of fat isoline), this would result in 18% non-compliance (y-axis). Similarly, Figure 23 shows the effect of a maximum limit on the mean concentration within a country if all products that do not comply with the limit are intercepted. In the above example, this would result in a reduction to 0.72 of the original mean. The results show that lowering a maximum limit becomes increasingly more effective in reducing the mean in the distribution when it moves towards the centre of the distribution. However, this would be achieved only when a correspondingly higher percentage of non-compliance is accepted.

FIGURE 22

TEQ, toxic equivalent

Figure 22. Theoretical effect of various maximum limits (see isolines in pg toxic equivalents per g of fat) on non-compliance of products, assuming a log-normal distribution with variable mean (x-axis) and a GSD of 2.0

FIGURE 22

TEQ, toxic equivalent

Figure 23. Theoretical effect of various maximum limits (see isolines in pg toxic equivalents per g of fat) on mean concentrations in truncated distributions, assuming an originally log-normal distribution with variable mean (x-axis) and a GSD of 2.0

Naturally, these outcomes depend on the within-food variation of concentrations in a certain food group and country. In Figures 22 and 23, a GSD of 2.0 was used (as estimated from the analysis of Hoogerbrugge et al., 2000). Figure 24 shows the relationship at various GSDs between non-compliance and the fraction of the mean remaining after all non-complying products (with concentrations higher than the maximum limit) have been removed. This graph demonstrates that, for log-normal distributions with very high GSDs, i.e. very wide, skewed distributions, lowering the mean is most effective. For example, for a GSD of 9.0, a reduction to 0.75 of the original mean will be achieved at the expense of 1% non-compliance, while for a GSD of 1.3, this reduction would cost 70% non-compliance.

FIGURE 22

In practice, different maximum limits are set for different food groups. Also, within a region, variation due to between-country differences must be considered in addition to within-food variation.

Figure 24. Theoretical relationship between the fraction of the mean concentrations remaining in truncated distributions (x-axis) and non-compliance (y-axis) for different GSDs (isolines), assuming an originally log-normal distribution

Four different scenarios were evaluated (see Table 18), which comprise a range of maximum limits positioned around those published by the Belgium Government (Belgisch Staatsblad, 1999). It should be noted that maximum limits for meat, eggs, fats and oils, and milk products are defined on a fat basis, and, in order to evaluate the effect of these maximum limits on intake, a conversion to a food basis is required. The mean fat contents of food groups were used for this purpose (Table 18). The four scenarios were applied to the national diets and to the western European and North American concentration distributions only, as national diets are believed to be the best approximation to real food consumption. Calculation of the regional concentration distributions requires an estimate of ‘between-country’ and ‘within-food’ variation in order to cover all components of variation. For an explanation of the between-country variation, see section 6.2.2. The ‘within-food’ variation was estimated from sets of individuals samples from the same area.

Table 18. Maximum limit scenarios for PCDDs/PCDFs

Food groupa

Unitsb

Scenario

 

 

 

 

 

1

2

3

4

Meat

pg/g fat

6.0

5.0

4.0

3.0

Eggs

pg/g fat

6.0

5.0

4.0

3.0

Fish

pg/g whole food

5.0

4.0

3.0

2.0

Fats and oils

pg/g fat

6.0

5.0

4.0

3.0

Dairy products

pg/g fat

6.0

5.0

4.0

3.0

Vegetable products

pg/g whole food

0.4

0.3

0.2

0.1

a

To convert to whole food basis, use the following fat contents: meat, 0.15 g/g; eggs, 0.12 g/g; fats and oils, 0.75 g/g; dairy products, 0.06 g/g

b

WHO toxic equivalents

Table 19 shows the results obtained by applying the scenarios presented in Table 18 for western Europe and North America. In both regions, the maximum limits considered would have the most radical effect on eggs. Other groups for which the maximum limits would have a relatively large effect are dairy products, meat (North America) and vegetable products. It should be stressed that the results in Table 19 are not truly representative of the situation in the regions considered. Much of the accuracy of the outcomes depends on the quality of the concentration distributions used in the exercise (see remarks in section 6.2.3).

Table 19. Effects of various maximum limit scenarios on non-compliance of products and fraction of mean concentrations of PCDD/PCDFs remaining

Source of data on concentration

Food group

Non-compliance (%)

Fraction of mean

1

2

3

4

1

2

3

4

North America

Meat

3.0

4.1

5.9

8.9

0.95

0.94

0.92

0.88

 

Eggs

7.1

9.2

12

17

0.90

0.88

0.85

0.80

 

Fish

0.8

1.3

2.2

4.5

0.98

0.98

0.96

0.93

 

Fats and oils

0.1

0.2

0.3

0.5

1.0

1.0

0.99

0.99

 

Dairy products

8.1

10

14

19

0.89

0.87

0.83

0.78

 

Vegetable products

1.3

2.2

4.4

12

0.98

0.96

0.93

0.85

Western Europe

Meat

1.1

1.6

2.4

3.9

0.98

0.97

0.96

0.94

 

Eggs

7.9

10

14

19

0.89

0.87

0.83

0.78

 

Fish

1.0

1.6

2.7

5.3

0.98

0.97

0.96

0.92

 

Fats and oils

0.1

0.2

0.3

0.5

1.0

1.0

0.99

0.99

 

Dairy products

3.6

4.8

6.8

10

0.95

0.93

0.91

0.87

 

Vegetable products

1.3

2.2

4.4

12

0.98

0.96

0.93

0.85

Table 19 shows that, given the variation in concentrations in the products, a significant reduction in the mean concentration could be made only when a large percentage of the products was removed from the market. Lowering the mean is more difficult for distributions that are less skewed (Figure 24). This would apply if the true GSD were found to be lower than that assumed for this analysis. On a national scale, e.g. during a local contamination incident in a particular food, the within-food GSD might temporarily be much higher than that expected from the GSD representing the usual variation in that food. Under such conditions, therefore, maximum limits would be very effective in reducing the average concentration in that food.

6.3.2 Dioxin intakes

The effects of various maximum limits on reducing the mean concentration in each food group result in an integral reduction in total intake. Table 20 shows the estimated reduction in intake when the four scenarios are applied to the national diets.

Table 20. Effect of various maximum limit scenarios on reducing the median intake of PCDDs/PCDFs in various diets

Scenario

Reduction in intake (%)

 

France

Netherlands

United Kingdom

USA

1

3.6

3.6

3.6

6.7

2

5.7

5.8

5.7

9.6

3

9.8

9.5

9.5

15

4

18

18

18

25

In concordance with the effect of the maximum limits on the mean concentrations in the truncated distribution (Table 19), the reduction in intake becomes increasingly more effective as the maximum limit is lowered. The somewhat larger reduction estimate for the North American diet is remarkable; it is due to the relatively larger effect of the maximum limits on the contamination of dairy products and meat in the diet in the USA.

The results shown in Tables 19 and 20 suggest that the relationship between the reduction in mean concentration and in intake is relatively straightforward. It should be stressed that this is due mainly to the assumption that between-country and within-food variations have universal values. If more accurate data were available on concentrations of these contaminants in foods for each food group and country throughout the world, the variation could be determined separately. Recalculating the effect of the scenarios would then possibly result in larger differences between countries.

6.4 Biomonitoring

6.4.1 Correlations between different matrices

PCDDs, PCDFs and coplanar PCBs can be determined in adipose tissue, whole blood or blood plasma or human milk to document human exposure, the values usually being reported on a lipid basis. Correlations have been demonstrated between the values found for TCDD and other PCDDs/PCDFs in serum and adipose tissue, in whole blood and adipose tissue and in whole blood and human milk (Patterson et al., 1988; Schecter et al., 1991; Päpke, 1998). The international toxic equivalents values for these pairs were similar, whereas those for the hepta- and octa congeners were somewhat different (Päpke, 1999). On a lipid basis, the ratio of serum or blood to tissue for TCDD content is approximately 1, and this ratio increases for higher chlorinated PCDDs/PCDFs (van den Berg et al., 1994). The lipid-based concentrations of all PCDD/PCDF congeners found in human milk, blood and adipose tissue are strikingly similar. The relative lipid content has been found to be the sole determinant of ‘adipose tissue:blood partition coefficients’ of highly lipophilic organic chemicals (i.e. chemicals with log n-octanol:water partition coefficients (PCo:w) > 4), indicating that, regardless of the identity and Po:w of these chemicals, their ‘adipose tissue:blood partition coefficient’ is equal to the ratio of lipid in adipose tissues and blood (Haddad et al., 2000).

The draft re-assessment of dioxins by the Environmental Protection Agency (2000a) in the USA also summarizes data on concentrations in human tissues. On the basis of those data, the assumption was made that the concentrations in human adipose tissue, blood and breast milk are similar on a lipid basis and that the concentrations in these tissues can be considered representative of overall body burden.

6.4.2 WHO-coordinated study of concentrations in human milk: Comparison of countries

The WHO Regional Office for Europe initiated a series of international studies on the concentrations of PCDDs, PCDFs and PCBs in human milk. The first round of studies took place in 1987–88, and the health risks of infants were re-assessed at a consultation in 1988 on the basis of the results. As the measured intakes found were the first to have been reported for many years and for many countries, the consultation recommended that the studies be repeated at 5-year intervals, in order to define the trends in intake. The second round of studies was performed in 1993, involving 19 countries in which samples could be collected and analytical data produced by the agreed deadline. Table 21 summarizes the results (WHO, 1996). Collection of samples for the third round began in 2000, the bulk of the samples being collected in 2001.

Table 21. Arithmetic average results of the second round of the WHO-coordinated studies on intake of PCDDs, PCDFs and coplanar PCBs from human milk (in pg/g of fat, expressed as toxic equivalents)

Country

Area

No. of samples per pool

Toxic equivalents

 

 

 

 

 

 

PCDDs/ PCDFs

non-ortho PCBs

2,3,3΄,4,4΄-PeCB+ 2,3΄,4,4΄,5-PeCB

Coplanar PCBs

Albania

Tirana

10

4.8

1.3

1.1

2.3

 

Librazhd

10

3.8

1.0

0.7

1.7

Austria

Vienna (urban)

13

11

8.3

3.4

12

 

Tulln (rural)

21

11

9.4

3.0

12

 

Brixlegg (industrial)

13

14

15

3.8

19

Belgium

Brabant

8

21

3.8

3.6

7.4

 

Liège

20

27

1.7

3.1

4.7

 

Brussels

6

27

4.0

3.9

7.8

Canada

Maritimes 1992

20

11

2.9

1.2–1.4

4.1–4.4

 

Québec 1992

20

13–14

5.1

1.7–1.9

6.8–7.0

 

Ontario 1992

20

18

5.8

1.8–2.0

7.7–7.9

 

Prairies 1992

20

15

2.3

0.9–1.1

3.2–3.4

 

British Columbia 1992

20

16

2.5

1.0–1.2

3.5–3.7

 

Canada (all provinces) 1992

100

14–15

3.8

1.5–1.7

5.3–5.5

 

Canada 1981

200

29

8.6

3.4–3.6

12

 

Gaspé

12

23

9.5

3.2–3.4

13

 

Basse Côte-Nord

4

15

20

5.7–6.0

25–26

 

Ungava Bay

4

14

9.8

4.3–4.6

14

 

Hudson Bay

5

21

13

8.0–8.3

21–22

Croatia

Krk

10

8.4

3.8

2.2

6.1

 

Zagreb

13

14

5.2

2.7

8.0

Czech Republic

Kladno

11

12

2.5

3.5

6.0

 

Uherske Hradiste

11

18

4.1

5.7

9.8

Denmark

Seven cities

48

15

2.3

2.2

4.5

Finland

Helsinki

10

22

1.9

2.7

4.6

 

Kuopio

24

12

1.0

1.4

2.4

Hungary

Budapest

20

8.5–8.6

0.8

0.8

1.7

 

Scentes

10

7.8

0.9

0.5

1.4

Lithuania

Palanga (coastal)

12

17

13

7.6

20

 

Anykshchiai (rural)

12

14

13

7.8

21

 

Vilnius (urban)

12

13

12

8.9

20

Netherlands

mean of 17 samples

17

22

8.8

2.5

11

Norway

Tromsø (coastal)

10

10

16

3.4

20

 

Hamar (rural)

10

9.3

7.4

3.0

10

 

Skien/Porsgrunn (industrial)

10

12–13

6.7

2.9

9.5

Pakistan

Lahore

14

3.9

1.9

0.4

2.3

Russian Federation

Arkhankelsk

1

15

2.9

5.7

8.6

 

Karthopol

1

5.9

2.0

2.9

4.9

Slovakia

Michalovce

10

15

6.4

7.0

13

 

Nitra

10

13

3.6

2.5

6.1

Spain

Bizkaia

19

19.4

6.7

3.9

11

 

Gipuzkoa

10

26

3.8

4.4

8.2

Ukraine

Kiev (1)

5

11

9.3

5.6

15

 

Kiev (2)

5

13

6.0

5.6

12

United Kingdom

Birmingham

20

18

2.5

1.8

4.3

 

Glasgow

23

15

2.6

1.3

4.0

In calculating toxic equivalents, data are shown for undetected values equated to 0 and to the limit of detection. If no difference was seen, a single value is presented. All figures rounded to two significant figures

To improve the reliability and comparability of the data from different laboratories, the WHO Regional Office also coordinated interlaboratory quality control studies. On the basis of the results, laboratories were considered qualified to analyse the collected human milk samples for PCDDs/PCDFs and PCBs. These WHO-coordinated studies are therefore an excellent basis for a reliable worldwide collection of data on intake.

Among the 19 countries that participated in the second round, the highest toxic equivalent (PCDD/PCDF) values (20–30 pg/g of lipid) were found in Belgium, Canada (Gaspé and Hudson Bay regions), Finland (Helsinki), The Netherlands and Spain (Gipuzkoa). In most of the human milk samples, the values were 10–20 pg/g of fat. The lowest values (4–10 pg/g of lipid) were measured in samples from Albania, Hungary, Pakistan and the less industrialized regions in Croatia, Norway and the Russian Federation.

Relatively high toxic equivalents of coplanar PCBs (20–30 pg/g of lipid) were found in two samples from Canada (Basse Côte-Nord, Hudson Bay) and in all three samples from Lithuania. One sample from Norway (coastal area) and one from Austria (industrial area near Brixlegg) also had relatively high concentrations of toxic equivalents of non-ortho PCBs (15–20 pg/g of lipid). It became apparent that the relative contributions of non-ortho and mono-ortho PCBs to the total toxic equivalents of coplanar PCBs differs from one region or country to another. Lower values were found in most human milk samples (< 15 pg/g). Although lower concentrations were clearly observed for Albania, Hungary and Pakistan, the ranking of countries for this group of compounds, from lower to higher levels, is in general not comparable with that for PCDD/PCDF concentrations.

In the second round of WHO studies, therefore, few regions and countries were identified in which the concentrations in human milk were different from those in other countries. The sample from the Hudson Bay region in Canada contained relatively high values for all compounds investigated, while the values were significantly lower for Albania, Hungary and Pakistan. The countries could not be ranked consistently with respect to the concentrations of the compounds analysed.

Regions could be identified in several countries in which the body burdens of PCDDs and PCDFs, coplanar PCBs or indicator PCBs were higher than in other areas and countries. Generally higher concentrations were observed for Belgium and The Netherlands for PCDDs and PCDFs and in Lithuania for non-ortho and mono-ortho PCBs. Exceptionally high concentrations of the six indicator PCBs were found in regions of Canada, the Czech Republic and Slovakia.

6.4.3 WHO-coordinated study: Time trends

As the third round of studies of intake is only now being conducted, time trends can be derived only from the first and second rounds and only for concentrations of PCDDs, PCDFs and marker PCBs, as other compounds were not determined in the first round.

The concentrations of PCDDs and PCDFs were not increasing. In fact, those in some countries tended to decrease, and dramatic decreases, by up to 50%, were seen in some countries in comparison with the 1987 study. The situation was less clear for PCBs, as different, and sometimes less reliable, analytical methods were used in many countries in the first of the two studies.

In an attempt to quantify the time trends in concentrations of dioxins in human milk, the overall annual decrease in Canada and Europe was estimated as 7.2% with a standard deviation of 0.8%. If such a decrease is assumed, the concentrations would be reduced to half their initial value within approximately 9.6 years.

6.4.4 Variation in dioxin content of breast milk samples

Data from a German dioxin reference testing programme (Bund/Länder-Arbeits-gruppe Dioxine, 2001) can be used to describe the variation in the dioxin toxic equivalent content of breast milk samples. As the number of samples is decreasing, data from 1995–98 (271 determinations) were used to determine the frequency distribution of the results for individual samples. As expected, a log-normal distribution was observed. Figure 25 shows the frequency distribution with a normal distribution fitted to the data, and Figure 26 shows the fitted log-normal distribution.

FIGURE 25

Figure 25. Frequency distribution of values for international dioxin toxic equivalent (I-TEq) content of breast milk samples from Germany, 1995-98, 271 determinations, fitted normal distribution

FIGURE 26

Figure 26. Frequency distribution of values for international dioxin toxic equivalent (I-TEq) content of breast milk samples from Germany, 1995-98, 271 determinations, fitted log-normal distribution

Table 22 presents the minimum, median, mean, maximum, 70th percentile, 90th percentile, 95th percentile and and 99th percentile values for dioxin toxic equivalents (Commission of the European Union, 2000a; Bund/Länder-Arbeitsgruppe Dioxine, 2001; Vieth, 2001). However, for the years 1996, 1997 and 1998, only 49, 40 and 47 determinations, respectively, were available for calculation of the percentiles. As percentiles exclude the effect of maximum contamination, the 99th percentile for those years is equal to the maximum. This value is therefore meaningless. Additionally, samples for 1996–98 and for the summarized range 1995–98 included pooled samples. As the frequency distribution was calculated without weighting, they represent the distribution of determinations and can be used only as an indication of the distribution. These percentiles should therefore be used with caution.

Table 22. Frequency distributions of toxic equivalents in German breast milk, 1995–98

Method for estimating toxic equivalents

Year

No. of determinations

Mean

Median

Minimum

Maximum

75th percentile

90th percentile

95th percentile

99th percentile

Individual

Pooled and individual

WHO

1995

135

 

21

19

6.2

 

27

32

38

46

 

1996

 

49

16

16

5.7

 

23

32

34

 

 

1997

 

40

14

14

7.0

 

17

20

26

 

 

1998

 

47

15

14

5.3

32

18

22

26

 

 

1995–98

271

18

17

5.3

46

23

30

34

44

 

International

1995

135

 

18

16

5.4

39

23

27

32

39

 

1996

 

49

14

14

4.9

30

20

27

30

 

 

1997

 

40

12

12

6.0

29

14

17

23

 

 

1998

 

47

13

12

4.7

29

16

20

23

 

 

1995–98

271

16

14

4.7

39

20

25

29

37

 

6.4.5 Dependence of toxic equivalents on age

Several groups have reported an effect of age on the body burden of PCDDs/PCDFs in adipose tissue, breast milk and blood. Päpke (1998) summarized the results of various authors, showing an increase of 0.4–0.8 pg/g of lipid per year of age, expressed as international toxic equivalents.

Age dependence was also evaluated as part of the German dioxin reference testing programme (Bund/Länder-Arbeitsgruppe Dioxine, 2001). Data from 1995–98 for breast milk samples from mothers aged 25–43 years who were nursing their first infant showed considerable variation by age (Figure 27). When linear regression was applied, a mean increase of 0.47 ng/kg of lipid per year was derived, expressed as toxic equivalents.

FIGURE 27

From Bund/Länder-Arbeitsgruppe Dioxine (2001)

Figure 27. Dependance on age of the international dioxin toxic equivalent (I-TEq) content of breast milk samples, Germany, 1995-98, 116 mothers nursing their first infant

6.4.6 Data from countries not participating in WHO field studies

As the third round of WHO-coordinated studies of intake of PCDDs, PCDFs and relevant PCB congeners is being conducted, data from other countries since 1993 were gathered from submissions or from the literature. Data from overviews were also used. Only information on background contamination is summarized here.

Brazil: The concentrations of PCDDs/PCDFs were determined in a pooled sample of breast milk from 40 mothers, 33 at their first lactation and seven at their second (age, 15–38 years), living in the urban area of Rio de Janeiro county in 1992. The dioxin content was 8.1 pg/g of lipid expressed as international toxic equivalents (Paumgartten et al., 2000).

Canada: Canada provided data on dioxin toxic equivalents in over 400 human milk samples collected across the country in 1986–87 (Ryan et al., 1993b). Of these, 100 samples, selected by province according to population, were analysed for PCDDs/PCDFs. The results showed the presence of 11 analytes with 2,3,7,8 substitution. The mean value for the country was about 15 pg/g of lipid. The data from the second round of the WHO-coordinated study showed a range of 11–29 pg/g of lipid and significant differences between the provinces. Additionally, in human milk collected in 1997 from residents of northern Canada (Keewatin), the mean content of PCDD/PCDF toxic equivalents was 4.9 pg/g of lipid, and that of non-ortho PCB toxic equivalents was 1.7 pg/g of lipid (Newsome & Ryan, 1999).

Egypt: Pooled breast milk samples were obtained from 45 mothers in Cairo, 30 from Ismaila and 12 from El-Menia, providing a representative overview for the country. The dioxin international toxic equivalents concentrations were 20–25 pg/g of lipid. Samples from Aswan, however, had a content of 11 pg/g of lipid. These findings corresponded to the finding of elevated concentrations of dioxin toxic equivalents in butter samples from northern Egypt, whereas butter samples from southern Egypt had the usual background contamination (Malisch et al., 2000).

European Union: The national average concentrations of PCDDs, PCDFs and coplanar PCBs (in pg of toxic equivalents per g of lipid) in representative human milk samples were summarized in the SCOOP report (Commission of the European Union, 2000a). The results are shown in Table 23.

Table 23. Toxic equivalents of PCDDs, PCDFs and coplanar PCBs in human milk (pg/g of fat) in countries of the European Union

Country

PCDDs/ PCDFs

Coplanar PCBs

 

Before 1990

1990–94

1995–99

Before 1990

1990–94

1995–99

Belgium

 

25

 

 

6.6

 

Denmark

18

17

 

 

18

 

Finland

20

13

 

25

12

 

France

 

 

16

 

 

 

Germany

31

21

14

 

 

 

Italy

25

 

 

 

 

 

Netherlands

34

24

 

 

21

 

Norway

 

10

 

 

29

 

Sweden

 

13

7.9

 

19

 

Range of means

18–34

10–25

8–16

25

7–29

No data

Before 1995, the national average concentrations ranged from 10 to 34 pg/g of lipid expressed as international toxic equivalents. During 1995–99, the national average concentrations ranged from 8 to 16 pg/g of lipid, some countries clearly showing a downwards trend.

The SCOOP report noted that the the mother’s intake throughout life, the number of previously breastfed children, residential factors, personal characteristics and various toxicokinetics factors result in the concentrations observed in human milk. Furthermore, several factors such as the methods of sample collection (e.g. time after childbirth, number of children), storage and transport of samples have been found to be relevant. All these factors may lead to the differences in mean concentrations of dioxins and PCBs observed in the various studies on human milk. Such differences can be even larger when the concentrations in individual samples from the same population are compared. The latest WHO field study on human milk samples revealed a coefficient of variation of 30–40% in the international toxic equivalent concentration in individual samples. Thus, both methodological factors and population and individual characteristics explain the differences seen in Table 23.

Finland: The dioxin concentrations in this population, who frequently eat fish from the Baltic Sea, are comparable to those seen in inhabitants of Seveso, Italy, after the accidental release of TCDD in 1976. High concentrations of international toxic equivalents were observed in frequent fish eaters (median, 170 pg/g of lipid; mean, 180; range, 51–420; n = 26). The value for men in Finland (median, 32 pg/g of lipid; mean, 33; range, 12–81; n = 45) was considered to be the normal concentration for men aged 40–70 years (Kiviranta et al., 2000).

France: A study was conducted in 1998 and 1999 on 244 breast milk samples from each of the eight French territorial zones. The mean concentration of PCDDs/PCDFs international toxic equivalents in milk was 17 pg/g of lipid (or 20 pg/g of lipid expressed as WHO toxic equivalents), with a range of 6.5–34 (7.8–41) (Fréry et al., 2000).

Germany: A summary of data from the State Institutes for Chemical Analysis of Food was evaluated for PCDD/PCDF international toxic equivalents in breast milk in Germany (Bund/Länder-Arbeitsgruppe Dioxine, 2001). Between 1986 and 1998, 1732 samples were collected from Baden-Württemberg, Bayern (Bavaria), Nordrhein-Westfalen and Niedersachsen (Lower Saxony) and the Federal Health Institute (Table 24). These data confirm a decrease from about 30 pg/g of lipid in 1990 to about 12 pg/g of lipid in 1998.

Table 24. Toxic equivalents of PCDDs/PCDFs in breast milk samples from Germany

Year

No. of samples

International toxic equivalents (pg/g fat)

WHO toxic equivalents (pg/g fat)

Minimum

Median

Mean

95th percentile

Maximum

Mean

Maximum

1986–90

728

5.6

29

31

 

87

36

 

1991

191

6.4

24

23

48

58

28

65

1992

171

3.5

21

21

39

48

24

54

1993

141

4.1

19

21

 

38

22

 

1994

90

4.9

18

17

37

44

20

50

1995

135

5.4

18

16

32

39

21

46

1996

81

4.9

14

14

30

30

16

35

1997

126

6.0

12

12

23

29

14

33

1998

69

4.7

13

12

23

29

15

32

Japan: A report on the tolerable daily intake of dioxins in Japan showed a decrease in the concentration of toxic equivalents of PCDDs/PCDFs in breast milk between 1973 and 1995, from about 30 pg/g of fat in the 1970s to about 20 pg/g of fat in the 1990s, and, for coplanar PCBs, from about 30 pg/g of fat to about 10 pg/g of fat (Environment Agency of Japan, 1999). A range of 12–37 pg/g of fat was reported for international toxic equivalents (Hashimoto et al., 1995; Hirakawa et al., 1995). The Environmental Protection Agency (2000a) in the USA reported that samples collected in Japan between 1989 and 1992 contained 28–45 pg/g of lipid expressed as WHO toxic equivalents of PCDDs/PCDFs.

Jordan: Thirty individual human milk samples were collected from five main towns in Jordan and pooled into six samples. The calculated international toxic equivalent values range from 0.19 ng/kg of lipid (excluding 14 undetectable congeners) and 97 ng/kg of lipid (Alawi et al., 1996).

New Zealand: New Zealand provided a press release that stated that the concentrations of dioxin in breast milk had fallen by about two-thirds during the past decade. The Environmental Protection Agency (2000a) in the USA reported a concentration of 21 pg/g of lipid for New Zealand in 1994, expressed as WHO toxic equivalents.

United Kingdom: Samples were obtained and pooled from 20 mothers in Birmingham and 23 mothers in Glasgow in 1993 and from 20 mothers in Cambridge in 1994. The concentrations of toxic equivalents of dioxins decreased from 29–37 ng/kg of milk fat in 1987–88 to 21–24 ng/kg of milk fat in 1993–94. The concentrations of PCB toxic equivalents in 1993 were 10–12 ng/kg of milk fat, giving a total of 31–34 ng/kg milk fat for dioxins and PCBs. PCBs were not determined in 1987–88 (Ministry of Agriculture, Fisheries and Food, 1997).

USA: The most extensive study of the body burdens of PCDDs/PCDFs is the National Human Adipose Tissue Survey (Environmental Protection Agency, 2000b), which involved analysis of 48 samples composited from 865 samples collected in 1987 from autopsied cadavers and patients undergoing surgery. The mean for all the samples was 28 pg/g of lipid expressed as international toxic equivalents or 33 pg/g of lipid expressed as WHO toxic equivalents. The tissue concentrations of international toxic equivalents were found to increase with age (0–14 years, 9.7 pg/g of lipid; 15–44 years 25 pg/g of lipid; > 45 years, 46 pg/g of lipid). In general, the average tissue concentrations were fairly uniform geographically. No significant difference was found on the basis of race or between males and females.

The Centers for Disease Control compiled data on blood concentrations in persons with no known unusual or site-specific exposure to PCDDs, PCDFs or coplanar PCB. The average concentration of toxic equivalents (calculated according to WHO TEFs from 1998, including dioxins and furans) was 20 pg/g of lipid. The coplanar PCBs 3,3΄,4,4΄-TCB, 3,4,4΄,5-TCB, 3,3΄,4,4΄,5-PeCB and 3,3΄,4,4΄,5,5΄-HxCB add about 2 pg/g of lipid expressed as PCB toxic equivalents. It was assumed that values for the missing PCBs would increase the estimate of PCB concentrations by another 3.3 pg/g of lipid, for a total value for PCB toxic equivalents of 5.3 pg/g of lipid and a total WHO toxic equivalent value (dioxins, furans and coplanar PCBs) of 25 pg/g of lipid. This is the lipid concentration of toxic equivalents assumed to repre-sent the background value in the USA (Environmental Protection Agency, 2000a).

The re-assessment of the Environmental Protection Agency (2000a) summarizes the concentrations of dioxins in human tissues in Europe, Japan and North America. In general, these data represent studies conducted in the late 1980s and early 1990s. The mean tissue concentrations of WHO toxic equivalents of PCDDs/PCDFs were 20–50 pg/g of lipid, with a mid-point value of 35 pg/g of lipid during that period. The mean for the USA was 33 pg/g of lipid, and the mean from the European and Japanese studies was 41 pg/g of lipid.

6.4.7 Pattern of congeners

As mentioned in section 6.4.4, the toxic equivalents of PCDDs/PCDFs and of PCBs vary from one region or country to another. The ranking of countries by contamination with PCBs is generally not comparable to that by PCDDs/PCDFs. Three countries were selected to illustrate the variation not only in toxic equivalents but also in congener patterns in breast milk samples obtained from the second round of the WHO study of intake (Tables 25 and 26). Librazhd, Albania, was selected to represent the areas with the least contamination, Germany to represent those with intermediate contamination and Canada (1981) to represent those with the heaviest contamination. In addition, serum samples collected in 1995–97 in the USA and analysed at the Centers for Disease Control and Prevention were included (with permission from the principal investigators). The samples were collected from 316 persons in six locations: Manchester, Missouri (n = 61), Times Beach, Missouri (n = 67), Jacksonville, Arkansas (n = 57), Oregon (n = 9), Wisconsin (n = 93) and North Carolina (n = 29). Table 25 shows the congener pattern on a weight basis, and Table 26 shows the congener pattern on the basis of WHO toxic equivalents. The contribution of each congener to the sum of the toxic equivalents is shown in Table 26.

Table 25. Concentrations of specific congeners of PCDDs, PCDFs and PCBs in human milk and serum (arithmetic mean in pg/g of lipid)

Compound

Breast milk

 

 

Serum (USA; n = 316)

Librazhd, Albania (n = 10; fat content, 4.7)

Berlin, Germany (n = 10; fat content, 5.0)

Canada (1981) (n = 200; fat content, 3.5)

PCDDs and PCDFs

 

 

 

 

2,3,7,8-TCDD

0.4

2.2

3.4

2.1

1,2,3,7,8-PeCDD

1.0

7.1

8.9

5.2

1,2,3,4,7,8-HxCDD

0.6

6.3

84

6.2

1,2,3,6,7,8-HxCDD

4.1

22

NR

73

1,2,3,7,8,9-HxCDD

1.1

5.1

18

7.1

1,2,3,4,6,7,8-HpCDD

6.3

23

94

79

1,2,3,4,6,7,8,9-OCDD

22

160

360

660

2,3,7,8-TCDF

0.3

ND (< 0.4)

4.2

0.7

1,2,3,7,8,-PeCDF

0.2

ND (< 0.4)

ND (< 1)

0.8

2,3,4,7,8-PeCDF

3.7

12

13

6.2

1,2,3,4,7,8-HxCDF

1.4

5.5

17

6.5

1,2,3,6,7,8-HxCDF

1.2

4.4

NR

5.3

1,2,3,7,8,9-HxCDF

ND (< 0.1)

ND. (< 0.4)

ND (< 1)

0.7

2,3,4,6,7,8-HxCDF

0.8

1.0

4.3

2.2

1,2,3,4,6,7,8-HpCDF

2.7

3.3

15

13

1,2,3,4,7,8,9-HpCDF

0.1

ND (< 1.0)

ND (< 1)

1.3

1,2,4,6,7,8,9-OCDF

0.3

ND (< 1.0)

ND (< 2)

2.1

Coplanar PCBs

3,3΄,4,4΄-TCB

0.6

3.0

ND (< 30)

31

3,4,4΄,5-TCB

NA

NA

NA

3.2

3,3΄,4,4΄,5-PeCB

9.7

83

83

18

3,3΄,4,4΄,5,5΄-HxCB

3.9

72

31

19

2,3,3΄,4,4΄-PeCB

1 300

8 000

ND (< 1800)

 

2,3΄,4,4΄,5-PeCB

5 600

19 000

34 000

 

2,3,4,4΄,5-PeCB

NA

NA

NA

NA

2,3΄,4,4΄,5΄-PeCB

NA

NA

NA

NA

2,3,3΄,4,4΄,5-HxCB

2 300

NA

10

NA

2,3,3΄,4,4΄,5΄-HxCB

ND (< 1000)

NA

ND (< 3000)

NA

2,3΄,4,4΄,5,5΄-HxCB

1 000

NA

NA

NA

2,3,3΄,4,4΄,5,5΄-HpCB

ND (< 500)

NA

NA

NA

Marker PCBs

2,4,4΄-TCB

ND (< 1500)

3 000

10 000

NA

2,2΄,5,5΄-TeCB

ND (< 1000)

1 000

1 700

NA

2,2΄,4,5,5΄-PeCB

ND (< 1000)

1 000

ND (< 900)

NA

2,2΄,3,4,4΄,5΄-HxCB

13 000

120 000

74 000

NA

2,2΄,4,4΄,5,5΄-HxCB

22 000

160 000

87 000

NA

2,2΄,3,4,4΄,5,5΄-HpCB

7 500

80 000

40 000

NA

Total

63 000

380 000

210 000

 

Other PCBs

2,3,4,4΄-TCB

ND (< 500)

NA

NA

NA

2,3΄,4,4΄-TCB

NA

NA

4 300

NA

2,4,4΄,5-TCB

2 000

NA

32 000

NA

2,3,3΄,4΄,6-PeCB

NA

NA

ND (< 1200)

NA

2,2΄,3,3΄,4,4΄,5-HpCB

NA

NA

24 000

NA

2,2΄,3,4,4΄,5΄,6-HpCB

NA

NA

5 500

NA

2,2΄3,4΄,5,5΄,6-HpCB

NA

NA

12 200

NA

NA, not analysed; ND, not detected; NR, not reported

Table 26. Concentrations of specific congeners of PCDDs, PCDFs and PCBs in human milk and serum (mean WHO equivalents, in pg/g of lipid; undetected = LOD/2)

Compound

Breast milk

Serum (USA; n = 316)

Librazhd, Albania
(n = 10)

Berlin, Germany
(n = 10)

Canada (1981)
(n = 200)

ng/g

%

ng/g

%

ng/g

%

ng/g

%

PCDDs

2,3,7,8-TCDD

0.4

6.1

2.2

8.7

3.4

8.0

2.1

11

1,2,3,7,8-PeCDD

0.1

1.5

0.71

2.8

0.89

2.1

0.52

2.7

1,2,3,4,7,8-HxCDD

0.06

0.9

0.63

2.5

8.4

20

0.62

3.3

1,2,3,6,7,8-HxCDD

0.41

6.3

2.16

8.5

NR

0.0

7.3

39

1,2,3,7,8,9-HxCDD

0.11

1.7

0.51

2.0

1.8

4.2

0.71

3.7

1,2,3,4,6,7,8-HpCDD

0.063

1.0

0.23

0.9

0.94

2.2

0.79

4.2

1,2,3,4,6,7,8,9-OCDD

0.0022

0.0

0.016

0.1

0.036

0.1

0.066

0.4

Sum PCDDs

1.1

18

6.5

26

16

36

12

64

PCDFs

2,3,7,8-TCDF

0.03

0.5

0.02

0.1

0.42

1.0

0.07

0.4

1,2,3,7,8,-PeCDF

0.01

0.2

0.01

0.0

0.025

0.1

0.04

0.2

2,3,4,7,8-PeCDF

1.9

28

6.0

24

6.5

15

3.1

16

1,2,3,4,7,8-HxCDF

0.14

2.1

0.55

2.2

1.7

4.0

0.65

3.4

1,2,3,6,7,8-HxCDF

0.12

1.8

0.44

1.7

NR

0.0

0.53

2.8

1,2,3,7,8,9-HxCDF

0.005

0.1

0.02

0.1

0.05

0.1

0.07

0.4

2,3,4,6,7,8-HxCDF

0.08

1.2

0.1

0.4

0.43

1.0

0.22

1.2

1,2,3,4,6,7,8-HpCDF

0.027

0.4

0.033

0.1

0.15

0.4

0.13

0.7

1,2,3,4,7,8,9-HpCDF

0.001

0.0

0.005

0.0

0.005

0.0

0.013

0.1

1,2,4,6,7,8,9-OCDF

0.00003

0.0

0.00005

0.0

0.0001

0.0

0.00021

0.0

Sum PCDFs

2.3

35

7.1

28

9.3

22

4.8

26

Sum PCDDs/PCDFs

3.4

52

14

54

25

58

17

89

Coplanar PCBs

3,3΄,4,4΄-TCB

0.00006

0.0

0.0003

0.0

0.0015

0.0

0.0031

0.0

3,4,4΄,5-TCB

NA

0.0

NA

0.0

NA

0.0

0.00032

0.0

3,3΄,4,4΄,5-PeCB

0.97

15

8.3

33

8.3

19

1.8

9.6

3,3΄,4,4΄,5,5΄-HxCB

0.039

0.6

0.72

2.8

0.31

0.7

0.19

1.0

2,3,3΄,4,4΄-PeCB

0.13

2.0

0.8

3.2

0.09

0.2

NA

0.0

2,3΄,4,4΄,5-PeCB

0.56

8.6

1.9

7.5

3.4

8.1

NA

0.0

2,3,4,4΄,5-PeCB

NA

NA

NA

0.0

NA

0.0

NA

0.0

2,3΄,4,4΄,5΄-PeCB

NA

NA

NA

0.0

NA

0.0

NA

0.0

2,3,3΄,4,4΄,5-HxCB

1.15

18

NA

0.0

5.0

12

NA

0.0

2,3,3΄,4,4΄,5΄-HxCB

0.25

3.8

NA

0.0

0.75

1.8

NA

0.0

2,3΄,4,4΄,5,5΄-HxCB

0.01

0.2

NA

0.0

NA

0.0

NA

0.0

2,3,3΄,4,4΄,5,5΄-HpCB

0.025

0.4

NA

0.0

NA

0.0

NA

0.0

Sum PCBs

3.1

48

12

46

18

42

2.0

11

Sum PCDDs, PCDFs and PCBs

6.5

100

25

100

43

100

19

100

NA, not analysed; NR, not reported

These variations cannot be generalized for whole countries or regions but give an indication of the variance. Dioxins appear to contribute imprtantly to the PCDD/PCDF toxic equivalents in the USA. It is difficult to compare the contribution of PCBs, as not all non-ortho and mono-ortho PCBs of interest were analysed; thus, comparisons of the sum of the PCB toxic equivalent values for the four selected countries are not appropriate. As an indication of the variation, however, the contributions of 3,3΄,4,4΄,5-PeCB, which was analysed in samples in all four countries, can be compared. The differences can be summarized by the factors shown in Table 27. Whereas the ratio of dioxins:furans in the four countries varied between 0.5 and 2.5, that of dioxins:3,3΄,4,4΄,5-PeCB varied between 0.8 and 6.7.

Table 27. Ratios of toxic equivalents for congeners in human milk and serum

Ratio of toxic equivalents

Breast milk

Serum (USA)

Librazhd, Albania

Berlin, Germany

Canada (1981)

Dioxins:furans

0.5

0.9

1.7

2.5

Dioxins:3,3΄,4,4΄,5-PeCB

1.2

0.8

1.9

6.7

6.4.8 Variation in body burden as an estimate of variation in intake

The integrating property of the long half-lives of PCDDs, PCDFs and coplanar PCBs suggests that the variation in body burdens among human populations should to some extent be comparable to the total variation in dietary intake of toxic equivalents of these compounds. A precise reflection of the variation is, however, unlikely, because, although the half-lives of the various congeners are long, they differ (see Table 2). These differences may slightly modulate the variation in intake of toxic equivalents, depending on the congener pattern.

The Environmental Protection Agency (2000a) of the USA presented an analysis of the variation in fat intake and in the body burdens of dioxin toxic equivalents, which suggests that the ratio of the 95th percentile to the average intake of dioxin toxic equivalents is < 2. Similar ratios were found for the 95th percentile to the average concentration of dioxin toxic equivalents in breast milk samples and serum in studies in the USA shown in Table 22. Fitting of log-normal distributions to these results of biomonitoring gave a GSD of approximately 1.5 (Table 28).

Table 28. Variation in WHO toxic equivalent concentrations in the results of biomonitoring

Data set

Year

Median

75th P

90th P

95th P

99th P

GSD

Breast milk, Germanya

1995–98

17.1

22.7

29.8

33.7

43.6

1.53

 

1995

19.1

26.6

32.2

37.8

45.6

1.53

Serum, USAb

1995–98

19.5

26.3

33.7

38.1

49.5

1.53

P, percentile

a Table 22

b Centers for Disease Control (2001)

Similarly, analysis of the mean dioxin toxic equivalents in breast milk from all the countries that participated in the WHO studies indicated that the variation in intake of dioxin toxic equivalents is also described by a GSD of < 2. Insertion of the data presented in Tables 21 and 29 into Eq. 1 from section 5.2 results in a GSD of 1.8.

Table 29. Between-country variation in mean WHO toxic equivalents in breast milk (ng/kg of lipid)

Country

Areaa

No. of individual samples

Average toxic equivalents

PCDDs/ PCDFs

Coplanar PCBs

Total

Pakistan

Lahore

14

3.90

2.30

6.20

Albania

 

20

4.30

2.00

6.30

Hungary

 

30

8.30

1.60

9.90

Russian Federation

 

2

10.55

6.75

17.30

Finland

 

34

14.79

3.05

17.84

Croatia

Weighted mean

23

11.28

7.17

18.46

Denmark

 

48

15.20

4.50

19.70

United Kingdom

 

43

16.46

4.14

20.60

Czech Republic

 

22

15.25

7.90

23.15

Slovakia

 

20

13.88

9.70

23.58

Norway

 

30

10.65

13.13

23.78

Ukraine

 

10

12.15

13.25

25.40

Austria

 

34

11.70

14.03

25.73

Germany

Berlin

10

16.55

11.70

28.25

Canada

 

425

21.52

9.19

30.72

Spain

 

29

21.50

9.77

31.28

Belgium

 

34

25.53

5.88

31.41

Netherlands

Mean for 17 persons

17

22.45

11.30

33.75

Lithuania

Weighted mean

36

14.77

20.53

35.30

Sum

 

881

 

 

 

a

Unless otherwise indicated, weighted mean for reported areas; areas and data shown in Table 21

The results shown in Tables 14 and 16 can be used to derive a GSD for the distributions of intake of 1.8 and 2.0 for the national diets and the GEMS/Food regional diets, respectively, which are somewhat higher than the GSDs observed for the breast milk samples.

6.5 Summary

Dietary intake of PCDDs, PCDFs and coplanar PCBs was assessed in a stochastic model that simulates the distribution of dietary intake of these compounds in the population, combining modelled log-normal distributions for concentrations in six food groups, log-normal distributions for total food consumption and the fraction that each food group contributes to food consumption. Concentration distributions were constructed from the data on occurrence submitted to the Committee. The distributions of food consumption were based on the results of national surveys of food consumption in various countries and the GEMS/Foods regional diets.

When the GEMS/Regional diets were used, the range of estimated intake of toxic equivalents of PCDDs and PCDFs was 13–140 pg/day at the median and 30–310 pg/day at the 90th percentile. The intakes of toxic equivalents of coplanar PCBs were 13–110 pg/day at the median and 37–300 pg/day at the 90th percentile. When intake estimates were based on national food consumption surveys, the concentra-tions were lower: for PCDDs and PCDFs, 66–84 pg/day at the median and 160–210 pg/day at the 90th percentile, and, for coplanar PCBs, 17–94 pg/day at the median and 50–250 pg/day at the 90th percentile. Estimates could not be made for the sum of PCDDs/PCDFs and PCBs, because data on concentrations were submitted separately by the participating countries.

Several uncertainties were identified in the dietary intake assessment: (i) The method used to account for undetectable concentrations in the aggregated data submitted was usually unknown. (ii) The data on occurrence consisted partly of the results of surveys, probably without random sampling, which would tend to result in an overrepresentation of high values. (iii) The GEMS/Food regional diets are based on data on food supply (apparent consumption), which are known to result in overestimates of food consumption by at least 15%. In general, these uncertainties suggest that the intakes given above are overestimates.

The concentrations of PCDDs, PCDFs and coplanar PCBs vary in products bought from retailers. In setting regulatory limits for coplanar compounds, the efficacy of such limits and possible undesired consequences of their use should be taken into account. The theoretical effects of various limits on reducing both intake and the fraction of products that comply with the limits were studied. In general, it was found that limits resulting in a substantial reduction of intake result in a concomitant increase in the number of products that do not comply with the limit.

The concentrations of PCDDs, PCDFs and coplanar PCBs in adipose tissue, whole blood, blood plasma or human milk provide an alternative measure of human exposure. Concentrations have been measured in breast milk in various national and international monitoring programmes. The results showed that the interindividual variation in concentration was comparable to the variation in simulated intake distributions.

7. PREVENTION AND CONTROL

7.1 Preharvest control

The best way to prevent contamination of crops with dioxin before harvest is to consider the possible sources and to avoid application of possibly contaminated products. As discussed in section 1.4.2, PCDDs/PCDFs are not produced intentionally but are found as unwanted and often unavoidable by-products in a number of industrial and thermal processes. Contaminated feeds are an important source for food of animal origin. The inventories of sources of emissions of dioxins into the environment should be extended to the domain of feed materials and feed production.

7.2 Decontamination

Food contaminated with coplanar PCBs is decontaminated only with difficulty. Defatting processes reduce the dioxin toxic equivalent content of food. For example, separation of cream from raw milk separates the contaminated milk fat from the uncontaminated whey. Refining could reduce the dioxin toxic equivalents in fat. For example, the high dioxin content of fish oil from fish in the Northern Hemisphere could be reduced considerably by refining, as deodorization by vacuum distillation also reduces the concentrations of lipophilic contaminants.

8. DOSE–RESPONSE ANALYSIS AND ESTIMATION OF RISK

This section presents an analysis of the dose–response relationships for human adverse health effects after exposure to TCDD and related compounds. The ideal result would be a description of an overall function, with attendant uncertainty, that would permit estimation of the probability that harm will occur after ingestion of a particular toxic equivalent dose of dioxin over some period of time. Regrettably, this ideal could not be attained with the information, knowledge and methods available to the Committee.

Instead, this section provides a conventional assessment of safety and an evaluation of three kinds of responses, with three methods of analysis, that together present a composite picture of the present state of scientific knowledge regarding the relation between intake of dioxins and risks. The safety assessment resulted in a recommendation for a provisional tolerable monthly intake (PTMI). The three assessments of dose–response relationships included an analysis of human cancer risk inferred from three epidemiological analyses of occupationally exposed persons, a semiquantitative analysis of severe chloracne and related effects drawn from epidemiological studies of two incidents of ingestion of contaminated cooking oil in Japan and Taiwan, and a ‘benchmark dose’ analysis based on numerous studies of non-cancer responses in animals. These assessments are presented to provide a context for the safety assessment and also to support mechanistic arguments about possible effects at low doses.

8.1 Safety assessment

The safety assessment was based on the results of tests in animals designed to identify the region of the dose–response function (for dioxin toxic equivalents alone, without consideration of the benefits of the food) that is presumed to result in no effect. This method produces a ‘tolerable intake’. It involves three steps: evaluation of the database to eliminate any studies that are of insufficient quality or would not predict the human response, identification of a single most sensitive study as the basis for establishing a safe level of intake, and expression of the results of this study according to a conventional algorithm, usually dividing the largest NOAEL by a safety factor with a default value of 100 but which can be assigned values of 3–1000. This procedure is designed to be biased towards health protection. Nevertheless, this protective bias is based on certain assumptions that are not always true, e.g. that human susceptibility to the chemical being assessed is within the range of that of the animals tested and that the groups of animals at the different doses in a test are large enough.

8.1.1 Pivotal toxicological data

Effects on the development of the reproductive system of male and female rat offspring exposed to TCDD in utero and during lactation as a result of low maternal doses during gestation have been reported in a number of studies (Table 30). Development of the immune system has also been reported to be affected at low maternal doses. The effects on the development of the reproductive system, particularly in male rat offspring of treated females, appear to be the most sensitive end-points.

Table 30. Studies of developmental effects in rat offspring after maternal exposure to TCDD

Reference

Study

Doses

Effects

Gehrs & Smialowicz (1999)

Fischer rats; single dose on GD 14

0
0.1
0.3
1.0


Suppressed DTH response in males
Suppressed DTH response in females

Murray et al. (1979)

Sprague-Dawley rats; three-generation study of reproductive toxicity; doses in ΅g/kg bw per day

0
0.001

0.01

0.1


No effect on reproductive end-points in F0, F1 or F2 generation
Diminished reproductive success in F1 and F2 generations
Frankly toxic to reproductive capability of F1 generation

Mably et al. (1992a)

Male Holtzman rat offspring; single dose
on GD 15

0
0.064
0.16
0.4


No effect
Decreased anogenital distance (absolute only; no change when normalized to body weight according to Birnbaum, 1998); delayed testicular descent; decreased seminal vesicle and ventral prostate weights; nonsignificant decreases in plasma testosterone and 5-dehydroxytestosterone concentrations; plasma luteinizing hormone concentration unaffected

Mably et al. (1992b)

Male Holtzman rat offspring; single dose on GD 15

0
0.064
0.16
0.4
1


No effect
Significantly increased latencies to mount, intromission and ejaculation; significantly decreased copulatory rates

Mably et al. (1992c)

Male Holtzman rat offspring; single dose on GD 15

0
0.064
0.16
0.4
1


Significantly decreased sperm parameters (dose-related) and testis and epididymis weights; no effect on reproductive success

Gray et al. (1997a)

Male Long-Evans rat offspring; single dose on GD 15

0
0.05

0.2
0.8


Altered eye opening; reduced ejaculated sperm counts
Delayed puberty; permanently reduced cauda epididymis sperm reserves

Gray et al. (1997b)

Female Long-Evans rat offspring; single dose on GD 15

0
0.05


0.2


0.8


No gross effects; histological effects on reproductive tract; no effects on reproductive success or on day of vaginal opening
Alterations in reproductive tract morphology; no effects on reproductive success or on day of vaginal opening
Alterations in reproductive tract morphology; delay in time to pregnancy and in vaginal opening (1 day)

Faqi et al. (1998)

Male Wistar rat offspring; loading and maintenance dose regimen before mating

0.025 ΅g/kg bw loading
0.005 ΅g/kg bw per week



0.06 ΅g/kg loading;
0.012 ΅g/kg bw per week

0.3 ΅g/kg loading;
0.06 ΅g/kg bw per week

Decreased daily sperm production, number of sperm, sperm transit rate; increased number of abnormal sperm; increased mounting and intromission latencies; no effect on serum testosterone concentration or on reproductive success
As above, but mounting and intromission latencies not significantly affected; no effect on serum testosterone concentrationor on reproductive success
As above, plus decreased serum testosterone concentration; no effect on reproductive success

Faqi & Chahoud (1998)

Female Wistar rat offspring; loading and maintenance dose regimen beginning 2 weeks before mating

0.025 ΅g/kg loading;
0.005 ΅g/kg bw per week
0.06 ΅g/kg loading;
0.012 ΅g/kg bw per week
0.3 ΅g/kg loading;
0.06 ΅g/kg bw per week

Delayed vaginal opening (< 1 day); no effect on reproductive organ weights or morphology

No effects


Delay in vaginal opening (< 1 day); decreased uterine weight; no effect on morphology

Ostby et al. (1999)

Long-Evans rats; 0.010 ΅g/kg bw per day for 90 days before mating

 

Accelerated eye opening; increased incidence of persistent vaginal thread in female offspring; no effect on sperm parameters in male offspring; no effect on number of pups, pup weight or sex ratio

Ohsako et al. (2001)

Male Holtzman rat offspring; single dose on GD15

0
0.0125



0.05





0.2


0.8


No effects on anogenital distance, testis, epididymis or ventral prostate weights, gonadotropin or testosterone concentration or sperm parameters
Decreased testicular weight (relative to body weight) only on day 120; decreased anogenital distance on day 120 (distance relative to body weight not reported); no effect on epididymis or prostate weights, hormone concentrations or sperm parameters
Decreased ventral prostate weight on day 120; decreased absolute anogenital distance; no effect on testicular weight or other parameters
Decreased ventral prostate weight on day 49; decreased absolute anogenital distance; no effect on testicular weight or other parameters

GD, gestation day; DTH, delayed-type hypersensitivity

Effects in the offspring of female rats treated with TCDD during gestation have been studied extensively (Mably et al. 1992a,b,c; Gray et al., 1997a,b; Faqi & Chahoud, 1998; Faqi et al., 1998; Gehrs & Smialowicz, 1999; Ostby et al., 1999; Ohsako et al., 2001). In most of these studies, effects were observed in male rat offspring at doses below those that induced effects in female offspring. In male rat offspring, the most sensitive reproductive end-points appeared to be effects on sperm counts and related parameters and effects on ventral prostate weight, depending on the study. In female rat offspring, the occurrence of a persistent thread of tissue across the vaginal opening (vaginal thread, see Flaws et al., 1997) was observed in some studies, but not others, at doses somewhat higher than those that induced effects in male rate offspring. However, the pattern of effects in male and female offspring was not completely consistent across studies and rat strains.

8.1.2 Adjustment from bolus dose to equivalent body burden in studies of developmental toxicity

The time of treatment in the studies of developmental toxicity, day 15 of gestation, marks the onset of the endocrine-sensitive phase of sexual differentiation in rats and therefore represents a critical time of fetal exposure. The bioavailability of TCDD to the fetus at a given maternal body burden may differ with a bolus dose and with dietary intake at steady state. As placental transfer is mediated via the blood, the serum rather than the tissue concentrations are critical in determining the extent of fetal exposure. After a bolus dose, the serum concentration of TCDD rises, before redistribution to the tissue compartments. In contrast, low-level long-term exposure does not significantly increase the serum concentration. This would suggest that the determinant of reproductive effects is the fetal concentration on day 15–16 of gestation, which, as noted above, is likely to be higher after a single bolus dose than after long-term intake of a lower concentration.

The issue of the difference in the fetal body burden after a single bolus dose and after low long-term exposure leading to a similar maternal body burden was addressed in a study in which radiolabel was measured in maternal and fetal tissues of Long-Evans rats on day 16 of gestation after short-term administration of [3H]TCDD. The rats were given [3H]TCDD by gavage at 1, 10 or 30 ng/kg bw in corn oil, 5 days per week, for 13 weeks. At the end of this period, the rats were mated, and dosing was continued daily throughout gestation. The regimen produced a steady-state concentration of TCDD in the dams; the average maternal and fetal body burdens on day 16 of gestation are shown in Table 31, where they are compared with the average maternal and fetal body burdens found on day 16 of gestation after a single dose of TCDD by gavage on day 15 of gestation.

Table 31. Average maternal and fetal body burdens on day 16 of gestation after a single dose and after short-term intake of TCDD by pregnant rats

Single dose on day 15 of gestationa

Short-term intakeb

Dose (ng/kg bw)

Body burden

Adjusted daily dose (ng/kg bw per day) c

Body burden

Maternal (ng/kg bw)

Fetal (ng/kg bw)

Maternal: fetal

Maternal (ng/kg bw)

Fetal (ng/kg bw)

Maternal: fetal

50

30

5.3

5.7

0.71

20

1.4

14

200

97

13

7.4

7.1

120

7.5

16

800

520

39

13

21

300

15

20

1000

580

56

10

 

 

 

 

a Data from Hurst et al. (2000a)

b Data from Hurst et al. (2000b)

c Adjusted to continuous intake from results for 5 days/week

Using the data of Hurst et al. (2000a,b) (Table 31), the Committee developed two alternative methods for calculating the maternal body burden after repeated dosing which would result in the same body burden in fetuses as after administration of a bolus dose on day 15 of gestation. In one method, a linear relationship is assumed between the administered dose and the maternal and fetal body burdens; in the other, a nonlinear relationship is assumed. The two methods and their justification are discussed below.

For the linear approach, least-squares linear fits of dose versus maternal and fetal body burdens were derived from the data in Table 31. As radiolabelled TCDD was used in both studies, a zero intercept was assumed for the fitted line. None of these fits showed what appeared to be any significant deviation from linearity (Figure 28). The fitted slopes were 0.609 [maternal body burden, bolus dose] 0.0534 [fetal body burden, bolus dose], 14.4 [maternal body burden, repeated dose] and 0.749 [fetal body burden, repeated dose]. Consequently, these data predict that the ratio of fetal:maternal body burden would be (0.0534/0.609)/(14.4/0.609) = 1.68 times higher after a bolus dose than after repeated dosing resulting in the same total dose.

FIGURE 28

Figure 28. Linear fits to the data of Hurst et al. (2000a,b) in Table 31

In general, TCDD displays non-linear kinetics, owing to its binding to inducible hepatic CYP proteins. This effect is seen in particular after repeated doses of TCDD, but only at doses that far exceed those used by Hurst et al. as repeated doses. In fact, clear linear accumulation of TCDD in the body is observed after long-term exposure to doses of TCDD ­ 125 ng/kg bw per day, which is sixfold higher than the maximum dose used by Hurst et al.. Similarly, in the study with a single dose, the time between administration of TCDD (day 15 of gestation) and determination of the body burden (day 16 of gestation) was too short to allow a significant effect of induction of hepatic proteins on the linear kinetics of TCDD. Consequently, linear kinetics are expected to obtain in this study also.

As further evidence for the non-linear approach, it was noted that extrapolation of the relationship between the fetal and maternal body burden did not intercept at zero as would have been expected, as radiolabelled TCDD was used in both studies. Therefore, a best-fit analysis was made of each data set in the range of fetal body burdens from 0 to 15 ng/kg bw, constraining the curves to pass through the origin. Both data sets were fit to power equations, with the following result:

single dose: (R2 = 0.999) (Hurst et al., 2000a)

short-term intake: R2 = 0.999) (Hurst et al., 2000b),

where Y is the maternal body burden of TCDD (ng/kg bw) and X is the fetal body burden (ng/kg bw).

These two equations were used to calculate the corresponding maternal body burdens for fetal body burdens ranging from 0 to 15.2 ng/kg bw (Table 32). The calculations indicated that the factor for converting maternal body burden after a single dose into a corresponding steady-state body burden is approximately 2.6. These do not provide assurance that the correct relationships have been found: they were used solely to express the data in such a way as to allow extrapolation between results obtained by giving single doses by gavage and in short-term studies with daily doses, for the purpose of estimating steady-state body burdens.

Table 32. Fetal and maternal body burdens of TCDD after single and short-term administration, on the basis of a non-linear relationship between dose and body burden

Fetal body burden (ng/kg bw)

Maternal body burden after single dose (ng/kg bw)

Steady-state maternal body burden after short-term intake (ng/kg bw)

Ratio of maternal body burden after single and short-term intake

1.2

5.0

12.3

2.5

1.4

5.9

14.6

2.5

1.7

7.5

18.6

2.5

1.8

8.0

20.0

2.5

1.9

8.5

21.0

2.5

2.1

10

25.0

2.5

3.0

15.5

39.0

2.5

5.3

31

78.6

2.5

6.3

38.5

99.0

2.6

7.5

47.5

122

2.6

8.0

52

134

2.6

9.0

60

156

2.6

13.2

95.7

251

2.6

15.2

113

299

2.7

8.1.3 Selection of critical studies

A number of biochemical changes, such as enzyme induction, altered expression of growth factors and enhanced oxidative stress, have been noted in experimental animals with TCDD body burdens within a lower range of 3–10 ng/kg bw. These biochemical effects were considered to be early markers of exposure to PCDDs, PCDFs and coplanar PCBs or events induced by these compounds in animals and in humans that may or may not result in adverse effects at higher body burdens.

The studies reviewed included those evaluated by IPCS in 1998 (Mably et al., 1992a,b,c; Rier et al., 1993; Gehrs et al., 1997; Gray et al., 1997a,b; Gehrs & Smailowicz, 1998) and two further studies (Faqi et al., 1998; Ohsako et al., 2001a,b). The most sensitive adverse effects reported were on development in male rat offspring and immunological deficits after prenatal exposure to TCDD (Table 33).

Table 33. Studies providing the body burdensa at which no effect and the lowest observed effect were seen for the most sensitive adverse effects of TCDD on developmental end-points in rats

Reference; rat strain

End-point

Dosing regimen

No-effect body burden (ng/kg bw)

Lowest effective body burden (ng/kg bw)

Ohsako et al. (2001); Holtzman

Ventral prostate weight; decreased anogenital distance in male offspring

Single oral bolus dose by gavage on day 15 of gestation

13

51

Faqi et al. (1998); Wistar

Decreased sperm production and altered sexual behaviour in male offspring

Loading dose; maintenance dose by subcutaneous injection

 

25

Gray et al. (1997a,b); Long-Evans

Accelerated eye opening and decreased sperm count in offspring

Single oral bolus dose by gavage on day 15 of gestation

 

28

Mably et al. (1992c); Holtzman

Decreased sperm count in offspring

Single oral bolus dose by gavage on day 15 of gestation

 

28

Gehrs et al (1997); Gerhs & Smialowicz (1998);

Immune suppression in offspring

Single oral bolus dose by gavage on day 14 of gestation

 

50

a Body burdens estimated froma linear fit to the data in Table 31

The reported increase in risk for endometriosis in rhesus monkeys after long-term intake of TCDD should be interpreted with caution, as the daily intake was not adequately reported. In addition, analyses conducted 13 years after the end of treatment showed high concentrations of coplanar PCBs in the blood of the monkeys with endometriosis, possibly from an unknown source of PCB. The Committee also noted that the doses used in some of the pivotal studies in rats (Table 33) would result in similar or lower equivalent human monthly intakes (EHMIs) than that from the LOEL for endometriosis in monkeys.

In one of the new studies (Ohsako et al., 2001), pregnant Holtzman rats were given a single oral dose of TCDD at 0–800 ng/kg bw on day 15 of gestation, and the male offspring were examined on day 49 or 120 after parturition. No changes were seen in testicular or epididymal weight nor in daily sperm production or sperm reserve at any dose. However, the weight of the urogenital complex, including the ventral prostate, was significantly reduced at doses of 200 and 800 ng/kg bw in rats killed on day 120. Moreover, the anogenital distance of male rats receiving doses > 50 ng/kg bw and killed at this time was significantly decreased. The Committee noted that administration of TCDD at any dose resulted in a dose-dependent increase in 5alpha-reductase type 2 mRNA and a decrease in androgen receptor mRNA in the ventral prostate of rats killed at day 49 but not in those killed at day 120, with no adverse sequelae at the lowest dose of 12.5 ng/kg bw. On the basis of 60% absorption and an assumption of a linear relationship for the data in Table 31, the equivalent maternal body burden after multiple doses at this NOEL would be 13 ng/kg bw. When the power model is used to fit the data in Table 31 (reported in Table 32), the lowest observed effective body burden was estimated to be 19.1 ng/kg bw. The LOEL of 50 ng/kg bw per day corresponds to an equivalent body burden of 51 ng/kg bw in the linear model and 76 ng/kg bw in the power model.

The lowest LOEL reported for effects on the reproductive system was in a study of male offspring of Wistar rats (Faqi et al., 1998). In that study, the dams were treated subcutaneously before mating and throughout mating, pregnancy and lactation. They received an initial loading dose of [14C]TCDD at 25, 60 or 300 ng/kg bw 2 weeks before mating and were then given a weekly maintenance dose of TCDD at 5, 12 or 60 ng/kg bw. The size of the maintenance doses was based on a reported elimination half-life of 3 weeks for adult rats. Effects on male reproduction were studied on postnatal days 70 and 170. The number of sperm per cauda epididymis was reduced in all TCDD-treated groups at puberty and at adulthood. Daily sperm production was permanently decreased, as was the sperm transit rate in the TCDD-exposed male offspring, thus increasing the time required by the sperm to pass through the cauda epididymis. Moreover, the male offspring of the TCDD-treated groups had increased numbers of abnormal sperm at adulthood. Mounting and intromission latencies were significantly increased for male offspring of dams at the lowest and highest doses but not for those at the intermediate dose. There was no clear dose–response relationship for most effects in the treated groups. Offspring of dams at the highest dose had decreased serum testosterone concentrations at adulthood and permanent changes in the testicular tubuli, including pyknotic nuclei and the occurrence of cell debris in the lumen. The fertility of the male offspring was not affected by any maternal dose.

In computing the dose required to produce the fetal concentration seen after the intial loading dose of 25 ng/kg bw, it was noted that this dose would have been reduced to 20 ng/kg bw before administration of the maintenance dose of 5 ng/kg bw on day 14. From the linear fit to the data in Table 32, the fetal body burden resulting from a maternal body burden of 20 ng/kg bw would be 1.04 ng/kg bw. The maintenance dose of 5 ng/kg bw administered on day 14 of gestation would make an additional contribution to the fetal body burden of 0.27 ng/kg bw, resulting in a total fetal body burden of 1.31 ng/kg bw. From the linear fit to the data in Table 32, a maternal TCDD body burden of 25 ng/kg bw at steady state would be needed to achieve this fetal body burden.

The studies summarized in Table 33 provide evidence of adverse effects on the reproductive system in the male (and female) offspring of pregnant rats exposed to TCDD. The studies demonstrate reductions in daily sperm production, cauda epididymidal sperm number and epididymis weight as well as accelerated eye opening, reduction in anogenital distance and feminized sexual behaviour of male offspring, associated with maternal steady-state body burdens of TCDD of > 25 ng/kg bw. Reductions in the weights of the testes and the size of sex accessory glands, such as the ventral prostate, in male offspring, development of external malformations of the genitalia in female offspring and reduced male and/or female fertility required higher maternal body burdens. The most sensitive end-points identified differed between studies. This might reflect strain differences in sensitivity and/or even minor differences in the experimental conditions, e.g. the diet. In one study, a single maternal dose of TCDD at 12.5 ng/kg bw by gavage decreased the androgen receptor mRNA level in the ventral prostate at puberty (postnatal day 49), indicating reduced androgenic responsiveness. However, at this dose, none of the above-mentioned adverse effects was seen in male offspring. The dose corresponds to an estimated maternal steady-state body burden of approximately 19 ng/kg bw. As for enzyme induction, altered expression of growth factors and enhanced oxidative stress, the Committee considered this effect to be an early marker of exposure to TCDD or an event that may or may not result in adverse effects at higher body burdens.

The studies listed in Table 33 were those considered by the Committee in identifying the LOEL and associated NOEL for assessment of tolerable intake. The LOEL was identified in the study of Faqi et al. (1998), and a NOEL was identified in the study of Ohsako et al. (2001). With the pharmacokinetic conversions described in Table 34, these two studies indicate a lowest effective maternal body burden of 25 ng/kg bw and a maternal body burden with no effect of 12.5 ng/kg bw.

8.1.4 Background body burdens in laboratory animals

As radiolabelled material was used in the studies considered for estimating body burden from the distribution of TCDD after multiple dosing, the known background concentrations of TCDD and other PCDDs and PCDFs in laboratory rodent tissues resulting from traces of these compounds in laboratory feed were ignored. The Committee identified two studies that could be used to predict the intake by rats of coplanar compounds in laboratory feed (van den Heuvel et al., 1994; Vilukesa et al., 1998a). The results of these studies were mutually consistent and predicted that ‘unexposed’ laboratory rats had body burdens of toxic equivalents of 3–12 ng/kg bw, depending on age. Thus, the maternal body burdens of TCDD in studies with radiolabelled material should be adjusted upwards by a minimum of 3 ng/kg bw to account for the background concentrations of unlabelled PCDDs and PCDFs. This may still lead to underestimates of the maternal body burden of toxic equivalents, as 3 ng/kg bw was the minimum dose in the two studies, and in one of the studies coplanar PCB compounds were not included.

Addition of 3 ng/kg bw to the LOEL identified by Faqi et al. (1998) and the NOEL identified by Ohsako et al. (2001) resulted in estimated total body burdens of toxic equivalents of 16 and 28 ng/kg bw, respectively.

8.1.5 Selection of safety factors

Safety factors typically used in establishing tolerable levels of intake for humans on the basis of results in animals usually include a factor for converting a LOEL to a NOEL (if needed), a factor for extrapolating from animals to humans, and factors to account for interindividual variations in susceptibility. Factors of 10 have often been used for interspecies extrapolation and human variability, and a factor of 3 or 10 for extrapolating from a LOEL to a NOEL (WHO, 1994).

As discussed above, a NOEL was identified for effects in male rat offspring; thus, no factor for conversion from a NOEL to a LOEL was needed for the EHMI derived from the study of Ohsako et al. (2001). As concluded at the 1998 WHO consultation (van Leeuwen & Younes, 2000), use of body burdens to scale doses from animal studies to equivalent human levels removes the need for safety factors for differences in toxicokinetics between animals and humans. A safety factor must be applied, however, to account for interindividual differences in toxicokinetics among humans. Only limited data were available on the toxicokinetics of TCDD in humans, and it was considered that the default factor of 3.2 (WHO, 1994) was appropriate.

Humans may be less sensitive than rats to some effects, but the conclusion is less certain for others, and it cannot be excluded that the most sensitive humans might be as sensitive to the adverse effects of TCDD as rats were in the pivotal studies. Therefore, it was concluded that no safety factor in either direction need be applied for differences in toxicodynamics among humans.

Use of the LOEL instead of the NOEL indicates the need for an additional safety factor (WHO, 1994). As the LOEL reported by Faqi et al. (1998) for the sensitive end-point was considered to be close to a NOEL and represented marginal effects, it was considered appropriate to apply a factor of 3 to account for use of a LOEL instead of a NOEL. This leads to an overall safety factor of 9.6 (3 x 3.2).

It was concluded that a total safety factor of 3.2 should be applied to the EHMI associated with the NOEL identified by Ohsako et al. (2001) and a total safety factor of 9.6 to the EHMI associated with the LOEL identified by Faqi et al. (1998).

8.1.6 Provisional tolerable monthly intake

The PTMI is a measure used by the Committee for food contaminants with cumulative properties. Its value represents the permitted human monthly exposure to contaminants unavoidably associated with otherwise wholesome and nutritious foods. As stated in the discussion of toxicokinetics, the long half-lives of PCDDs, PCDFs and coplanar PCBs result in only a small or even negligible effect of each daily ingestion on overall intake. The appropriate period over which to average intake of dioxin is therefore months. Only after consideration of the total or average intake of PCDDs, PCDFs and coplanar PCBs over months can the long-term or short-term risk to health of intake of these compounds be estimated reliably. The tolerable intake should therefore be assessed over 1 month or longer. To encourage this view, the tolerable intake is expressed as a monthly value.

As shown in Table 34, use of the linear model to extrapolate the maternal body burden at the NOEL after a single dose in the study of Ohsako et al. (2001) to that expected after multiple doses shows that the EHMI expected to result in a body burden that is below that which had effects in animals is 237 pg/kg bw. The PTMI derived by application of the safety factor of 3.2 to this EHMI is 74 pg/kg bw.

Table 34. Summary of four calculations of provisional tolerable monthly intakes (PTMIs) for PCDDs, PCDFs and coplanar PCBs

 

Linear model

 

Power model

 

 

Ohsako et al.

Faqi et al.

Ohsako et al.

Faqi et al.

 

(2001)

(1998)

(2001)

(1998)

Administered dose (ng/kg bw)a

12.5

 

12.5

 

Maternal body burden (ng/kg bw)b

7.6

25b

7.6

25b

Equivalent maternal body burden with repeated dosing (ng/kg bw)

13c

25c

19d

39d

Body burden from feed (ng/kg bw)

3

3

3

3

Total body burden (ng/kg bw)

16

28

22

42

EHMI (pg/kg bw per month)

237

423

330

630

Safety factor

3.2

9.6

3.2

9.6

PTMI (pg/kg bw per month)

74

44

103

66

EHMI, equivalent human monthly intake

a Bolus dose (NOEL)

b Target maternal body burden after repeated dosing (LOEL)

c Linear relationship between fetal and maternal body burden assumed from data in Table 30

d Non-linear relationship between fetal and maternal body burden assumed from data in Table 30

e For humans, 7.6 year half-life and 50% uptake from food assumed (Eq. 1)

Similarly, as shown in Table 34, the PTMI derived by application of the safety factor of 9.6 to the EHMI derived from the study by Faqi et al. (1998) is 44 pg/kg bw.

As also shown in Table 34, use of the power model to extrapolate the maternal body burden incurred after a single dose to that after multiple doses results in PTMIs of 103 pg/kg bw for the NOEL of Ohsako et al. (2001) and 66 pg/kg bw for the LOEL of Faqi et al. (1998).

The range of PTMIs derived from the two studies, with either the linear or the power model for extrapolating the maternal body burden after single or multiple doses, is 40–100 pg/kg bw per month. The mid-point of this range, 70 pg/kg bw per month, was chosen as the PTMI for use in the safety assessment. Further-more, in concordance with the 1998 WHO consultation (van Leeuwen & Younes, 2000), the Committee concluded that this tolerable intake should be applied to the intake of PCDDs, PCDFs and coplanar compounds expressed in WHO TEFs.

8.2 Dose–response relationship above the tolerable intake

Effects due to ingestion of dioxins follow from binding of these compounds to the Ah receptor and subsequent biological events. Different animal species have different reactions, both qualitatively and quantitatively, although a few responses, such as mixed-function oxidase enzyme induction, seem to be near-universal.

8.2.1 Effects of uncertainty in estimates of intake

Until recently, researchers assumed that clearance of TCDD and its relatives was (or could be approximated by) a simple first-order process of the familiar kind. In this method, the change in concentration over time is used to estimate a half-life for depuration (i.e., the time required to reduce the concentration by half), from which a rate constant can be calculated. The model is not consistent with several observations, however, most notably measurements of clearance rates at high doses in the Yusho and Yu-Cheng rice oil poisoning incidents. It is also not consistent with the observation that background concentrations of dioxin are excreted both unchanged and as an oxidized metabolite (not identified in humans) (Wendling et al., 1990) Excretion unchanged is a diffusion-limited process, while excretion of a metabolite is probably a first-order reaction (assuming that oxidation is rate-limiting); however, the data suggest that considering these two processes as one does not introduce a large error. These observations show that, for body burdens > 1000 ng/kg bw, the initial depuration half-life is about 1 year, while those based on measure-ments of blood concentrations nearer to the background level are about 10 years. Back-calculation of the initial doses received in the Yusho and Yu-Cheng incidents from concentrations measured many years later results in underestimates of the actual doses by about an order of magnitude. Carrier et al. (1995a,b), Andersen and co-workers (Leung et al., 1990a,b) and Zeilmaker and van Eijkeren (1997, 1998; Zeilmaker et al., 1999) have shown that more complex models are required to account for the clearance rates of high doses. They have explained this discrepancy as a function of variable storage of dioxins in proteins induced in the liver. Depuration from this store is much more rapid than that from lipid stores, and, in addition, induction oxidase enzymes may further speed clearance.

In general, estimates of intake generated by back-calculating from a blood-lipid concentration measured years after peak intake are uncertain because they assume that depuration from peak intake follows first-order kinetics. Thus, estimates are affected by a systematic error; the Committee estimated that the highest lipid concentrations were actually about 50% higher. Correction for this systematic error would tend to reduce the slopes of the curves fitted to data based on first-order kinetics.

In most of the incidents of occupational exposure and in Seveso, some of the exposure occurred via dermal absorption. As TCDD and compounds with similar chemical properties dissolve in and are strongly bound to the skin (Banks & Birnbaum, 1991; Weber et al., 1991), estimates of local skin dose based on analyses of blood or internal organs will be systematically low. The Committee commented on instances in which such uncertainty should be taken into account.

8.2.2 Quantitative uncertainty in human cancer risk assessment

As noted in the review of epidemiological studies of cancer, several authors found dose-related trends in the risks for cancers at all sites and for cancers at various specific sites. To explore the potential quantitative range of cancer risks suggested by these studies, a meta-analysis was conducted by the Committee of the combined data on all cancers from the epidemiological studies for which appropriate data were available. Three such studies were identified: the study of German pesticide and herbicide workers by Flesch-Janys (1998), the study by Ott & Zober (1996) of German workers whose exposure resulted from an industrial accident and the updated follow-up by Steenland et  al. (1999, 2001) of a cohort of workers exposed to dioxin-contaminated chemicals at several plants in the USA.

In order to compare risks indicated by different studies, exposure must be quantified by a common metric that is plausibly related to risk. As the analysis was based on published data, selection of the dose metric was limited by the way in which the data were reported. Cumulative serum lipid concentration was selected for the meta-analysis. This measure is a form of cumulative exposure, with serum lipid concentration representing a measure of internal exposure.

Flesch-Janys (1998) categorized observed and expected numbers of cancer deaths by quartiles of cumulative concentration of TCCD toxic equivalents in serum lipid, reduced by the cumulative TCDD toxic equivalents contributed by the background. Since only ranges were provided by Flesch-Janys (1998), average values within these ranges were assumed (the mid-point for bounded ranges and twice the lower bound for the highest (unbounded) ranges). The observed and expected numbers of cancer deaths, relative risks and estimated cumulative concentrations are shown in Table 35.

Table 35. Dose–response relationships found in three epidemiological studies of persons exposed to dioxins

Reference

Cumulative concentration of TCDD toxic equivalents

No. of deaths from cancer in serum (ng/kg–years)

Observed

Expected

SMR

Flesch-Janys et al. (1998)

180

25

23.3

107

 

988

34

20.8

164

 

3 416

31

23.3

133

 

10 425

34

20.8

164

Ott & Zober (1996)

2 020

8

10.0

80

 

22 218

8

6.7

120

 

60 593

8

5.7

140

 

161 582

7

3.5

200

Steenland et al. (1999, 2000)

168

67

68.4

98

 

428

27

30.0

90

 

866

31

27.2

114

 

3 738

30

25.4

118

 

5 232

34

25.6

133

 

14 011

33

19.5

169

 

60 000

34

20.6

165

SMR, standardized mortality ratio

Ott & Zober (1996) categorized the deaths from cancer and SMRs by the estimated total dose of TCDD (΅g/kg bw). To convert these total doses to cumulative concentrations of TCDD in serum lipid, they were divided by 0.25 (assuming an average per cent body fat of 25%) and then divided by the decay rate (0.099/year) corresponding to a half-life of 7 years. As exposure in this study was the result of an accident and follow-up continued for 40 years after the accident, this was considered to be an adequate approximation, given the other uncertainties in exposure. The average exposures within categories were estimated as described above. The results are shown in Table 35.

Steenland et al. (1999) categorized observed and expected cancer deaths (the latter based on age-, sex- and calendar year-adjusted mortality rates in the USA) by septiles of a cumulative exposure score defined such that exposures during the most recent 15 years did not contribute to the score (i.e. exposure ‘lagged’ by 15 years). Steenland et  al. (2001) computed risk ratios categorized by septiles of cumulative TCDD concentration in serum lipid, including the contribution of background concentrations, with the same 15-year lag. In these risk ratios, the group with low exposure was used as the reference group. They are therefore not suitable for use in the meta-analysis. However, as a strong correlation was found between the cumulative exposure score and cumulative serum concentration (Spearman correlation, 0.9; Steenland et al., 2001), it is reasonable to use the septiles of cumulative exposure estimated by Steenland et al. (2001) as the exposures in the septiles defined by the exposure index (Steenland et al., 1999). The average exposure in each septile category was estimated by the approach used for the data of Flesch-Janys (1998), as described above. The results are shown in Table 35.

Several differences can be seen in the estimates obtained from these three studies. First, Steenland et al. estimated total exposure including background, whereas the other two groups estimated the exposure additional to background. Second, Steenland et al. used a 15-year lag in estimating exposure, whereas the other groups did not. These two differences should offset each other to some extent. Given that follow-up of the cohort studied by Steenland et al. in most cases extended for many years after the time at which exposure was greatest, cumulative exposure that lagged by 15 years should not differ greatly from unlagged exposure. Further, as background exposure is generally lower than occupational exposure, adjustment for background exposure should change the estimates only minimally. For these reasons, the differences in expression of exposure were considered to be of no major consequence, and no adjustments were made to account for them. Finally, the estimate of Flesch-Janys included total TCDD toxic equivalents, whereas Ott & Zober and Steenland et al. quantified only TCDD. This would result in an underestimate of total toxic equivalents in the last cohort, because some members were involved in production of chemicals that give rise to other PCDD/PCDF congeners; members of the cohort of Ott & Zober were exposed essentially only to TCDD.

In addition to other uncertainties in the exposure assessment, there is evidence that, at higher concentrations, liver enzymes are induced that increase the elimination rate of TCDD-like compounds. This effect was not accounted for in any of the three assessments of exposure used in the meta-analysis, but rather a first-order elimination process was assumed in each case. On the basis of the estimated maximum body burdens in these studies, the degree of underestimation of cumulative exposures due to failure to account for enzyme induction would be expected to be at most a factor of 1.5 (Zeilmaker et al., 1998; van der Molen et al., 2000).

FIGURE 29

Circles, from Flesch-Janys et al. (1998a); triangles, from Ott & Zober (1996); no symbol, from Steenland et al. (1999, 2001); line, model prediction

Figure 29. Standardized mortality ratios (SMRs) with 95% confidence bounds from three studies in which participants were classified by cumulative lipid concentration of TCDD (Table 35), including fit of linear model

Figure 29 shows the SMRs and corresponding 95% confidence intervals plotted against cumulative TCDD concentration in serum lipid. A log scale was used on the axis representing exposure in order to distinguish the different exposure groups more clearly. Likelihood ratio tests for a linear trend in these data were statistically significant, both when the background SMR was set equal to 100 (p < 0.0001) and when the background SMR was estimated (p = 0.01). The Committee concluded that these data provide statistical evidence for an effect of dioxin on the risk for human cancer.

This analysis did not address the likelihood that confounding by lifestyle factors or occupational exposure to other chemicals may have been responsible for the observed responses in the individual studies. However, the trend analysis with estimation of the background SMR was effectively a comparison of the responses of workers exposed to different concentrations of TCDD. As these workers had similar jobs and presumably similar socioeconomic status, confounding as an explanation for the associations should be less of a concern than in direct comparisons of exposed workers with external comparison groups.

Despite the statistical significance of the trend tests, the data are only marginally consistent, as evaluated with a goodness-of-fit test, with no effect of exposure and a background SMR of 124 (goodness-of-fit p value of 0.08). However, a simple goodness-of-fit test does not take into account the fact that the alternative hypothesis is a trend of increasing response with increasing exposure to TCDD. A trend test is more powerful for this alternative and accordingly is more appropriate. Consequently, in this instance the results of the trend test should take precedence over those of a goodness-of-fit test.

A series of trend tests were applied to the data in Table 35 to determine the lowest concentration at which there was statistically significant evidence for an effect, and the highest concentration at which there was no such evidence. In this procedure, the data were ordered with respect to cumulative serum concentration, and a test for linear trend (with the background SMR estimated) was applied. Then, the data for the highest concentration were omitted and the trend test was applied to the remaining data. This procedure was performed repeatedly until it was evident that statistical significance would not be obtained if additional data were omitted. The concentration below the lowest dose at which statistical significance was found can be interpreted in the same way as a NOEL derived from data on experimental animals: as the highest concentration at which a statistically significant effect of exposure is not seen.

Table 36 gives the results of the trend analysis. As noted above, the trend for all the data was significant (p = 0.010). When the data at the highest concentration were omitted, the trend became more significant (p = 0.008) and the slope of the trend increased. As successive data points were omitted, the trend remained significant (p < 0.05), until only the data corresponding to a cumulative exposure of 5232 ng/kg–years were left. However, as additional data points were omitted, the trend again became significant (p = 0.05) when the highest concentration remaining was 988 ng/kg–years. The highest concentration below this effect level is 866 ng/kg–years. Thus, there is convincing statistical evidence of an effect of concentrations > 5232 ng/kg–years and marginal evidence for an effect of concentrations > 866 ng/kg–years. On the basis of the rate at which dioxin accumulates in the body (assuming 50% uptake rate, a half-life of 7.6 years and sequestration of TCDD toxic equivalents in lipid, which comprises 25% of body weight), the daily intake required for a cumulative lifetime (to age 70) concentration of 5232 ng/kg–years is 44 pg/kg bw per day, and the estimated daily intake necessary for a cumulative concentration of 866 ng/kg–years is 7.3 pg/kg bw.

Table 36. Results of test for dose–response trend applied to a given exposed group and all groups with lower exposure to TCDD toxic equivalents

p for trend

Linear slope (ng/kg– years)–1

Cumulative lipid concentration (ng/kg–years)

Deaths from cancer

Reference

Observed

Expected

SMR

0.010

6.3 Χ 10–6

161 582

7

3.5

200

Ott & Zober (1996)

0.016

8.3 Χ 10–6

60 593

8

5.7

140

Ott & Zober (1996)

0.013

1.0 Χ 10–5

60 000

34

20.6

165

Steenland et al. (1999)

0.0080

3.5 Χ 10–5

22 218

8

6.7

120

Ott & Zober (1996)

0.0018

5.0 Χ 10–5

14 011

33

19.5

169

Steenland et al. (1999)

0.013

5.7 Χ 10–5

10 425

34

20.8

164

Flesch-Janys (1998)

0.12

5.8 Χ 10–5

5 232

34

25.6

133

Steenland et al. (1999)

0.095

6.4 Χ 10–5

3 738

30

25.4

118

Steenland et al. (1999)

0.18

9.7 Χ 10–5

3 416

31

23.3

133

Flesch-Janys (1998)

0.45

1.3 Χ 10–4

2 020

8

10.0

80

Ott & Zober (1996)

0.050

5.5 Χ 10–4

988

34

20.8

164

Flesch-Janys (1998)

0.61

1.6 Χ 10–4

866

31

27.2

114

Steenland et al. (1999)

NS

 

428

27

30.0

90

Steenland et al. (1999)

NS

 

180

25

23.3

107

Flesch-Janys (1998)

NS

 

168

67

68.4

98

Steenland et al. (1999)

To further explore the increase in risk for death from cancer potentially resulting from exposure to TCDD toxic equivalents, consistent with the data in Table 35, a dose–response model for SMR was fit to these data. In this model, it is assumed that the SMR depends linearly on the cumulative serum lipid concentration (in ng/kg–year) of TCDD,

SMR = 100 Χ alpha Χ (1 + beta Χ CSLC),

Eq. 1

where 100 Χ alpha is the baseline SMR and beta is the linear parameter that gauges the potential carcinogenic potency of TCDD. A linear model for SMR was considered adequate for this analysis because the model fit the data on dose–response in Table 35 adequately (as will be shown below), and the model is not proposed for extrapolation outside the range of the data. This model was fit both with the baseline SMR fixed at 100 (alpha = 1) and with a variable baseline SMR (estimated alpha).

Table 37 summarizes the results of fitting the model to the data in Table 35. The hypothesis that the baseline SMR = 100 was appropriate (alpha = 1) was rejected (p = 0.01). Consequently, the model in which the baseline SMR is estimated was preferred. This model provided an adequate fit to the data (goodness-of-fit p = 0.37), produced a baseline SMR estimate of 100 Χ alpha = 117 (90% CI:1.06, 1.28), and predicted that each ng/kg–year of cumulative lipid concentration increased the relative risk by alpha = 6.89 Χ 10–6 (90% CI, 2.17 Χ 10–6, 1.28 Χ 10–5). The best-fitting model with SMR = 117 is shown in Figure 29.

Table 37. Results of fitting the linear relative risk model to the data in Table 34, with corresponding estimates of ED10 (continuous daily intake throughout life in pg/kg bw per day estimated to result in an additional lifetime risk for death from cancer of 0.1), ED05 and ED01 (90% confidence intervals in parentheses)

SMR

Potency
(ng/kg–years)–1

Goodness-of-fit p value

ED10
(pg/kg bw per day)

ED05
(pg/kg bw per day)

ED01
(pg/kg bw per day)

100

1.19 Χ 10–5

0.08

251

122

24

 

(6.7 Χ 10–5, 1.8 Χ 10–6)

 

(167. 449)

(81, 218)

(16, 43)

Variablea

6.89 Χ 10–6

0.37

435

212

41

 

(2.2 Χ 10–6, 1.3 Χ 10–5)

 

(234, 1379)

(114, 670)

(22, 131)

SMR, standardized mortality ratio; ED, effective dose

a Baseline SMR (100alpha), 117 (106, 128)

The results obtained with these model fits were converted to estimates of ED10, ED05 and ED01. An ED10, for example, is defined as the average constant lifetime intake of dioxin toxic equivalents (pg/day) that corresponds to an increase by 0.1 in the lifetime risk for death from cancer due to intake of dioxin (a 10% additional lifetime risk).

In making these calculations, a constant daily intake of TCDD toxic equivalents was assumed. The corresponding cumulative lipid concentration (ng/kg–years) was computed as a function of age, assuming a first-order elimination process with a 7.6-year half-life, a 50% uptake of coplanar compounds from food, and sequestration of essentially all the TCDD toxic equivalents in lipid, which comprises 25% of body weight. For each 5-year age interval, the corresponding average cumulative intake of TCDD was calculated, and the corresponding age-specific baseline cancer mortality rate was modified by multiplying it by the relative risk predicted from the model. The probability of survival to a given age was calculated from the baseline rates of mortality from all causes, also suitably modified to account for the relative risk for cancer predicted from the model. The probability of dying from cancer in a given 5-year age interval was calculated as the probability of surviving to the beginning of the interval, multiplied times the annual mortality rate for that age interval (modified in accordance with the dose–response model to account for intake of dioxin), multiplied by 5 years (the width of the age interval). The total lifetime probability of dying from cancer was then computed by summing the contributions from each 5-year age interval. The corresponding lifetime probability of dying from cancer, assuming no exposure to dioxin (in reality, no additional exposure over background) was calculated in the same manner. The difference between the lifetime risk with exposure to dioxin and the risk with no exposure is defined as the estimate of the ‘additional’ lifetime risk for cancer. To calculate an ED10, the daily intake was adjusted to make this additional risk equal to 0.1.

The mortality rates used in this calculation were those for both sexes and all races combined in the USA for the years 1985–90. As these rates already include any contribution to cancer mortality from background exposure to dioxin (and the exposure used in the modeling can reasonably be considered to be estimates of exposure above background), the additional risk obtained in this calculation is best interpreted as additional risk from exposure above background levels of exposure.

The exposures were lagged by 15 years in the study of Steenland et al. (1999) and not lagged in the other two studies. In the calculations described above, a 15-year lag was used. Some lag would appear to be appropriate, as exposure immediately before death is not likely to affect the cancer response. As Steenland et al. (1999) used a 15-year lag, and their study comprised most of the data, the Committee decided to use a 15-year lag in the calculations. Not incorporating a lag would increase the estimated risks by roughly 40%.

The ED01, ED05 and ED10 determined by this method are shown in Table 37. The model with the baseline variable was consistent with the data (goodness-of-fit p value = 0.36) and predicted an ED10 of 435 pg/kg bw per day (90% CI, 234–1379), an ED05 of 212 pg/kg bw per day (90% CI, 114–670) and an ED01 of 41 pg/kg bw per day (90% CI, 22–131). These values are best interpreted as concentrations above the background concentrations of TCDD toxic equivalents that are predicted to increase the risk for death from cancer above the background rate, by the given additional amount. The background risk for death from cancer includes that due to background TCDD.

In its draft document on the health effects of dioxin, the Environmental Protection Agency in the USA applied a linear model to data for the same three cohorts to predict the lifetime risk for cancer. This analysis differed from that conducted by the Committee in three main ways. First, rather than using the data of Steenland et al. (1999, 2001), they used data from Fingerhut et al. (1991), which covered the same cohort but included five fewer years of follow-up and a less detailed assessment of exposure. Secondly, the Environmental Protection Agency used average body burden as the exposure metric, whereas the Committee used cumulative serum concentration. Thirdly, the Environmental Protection Agency assumed a baseline SMR of 100, whereas the Committee allowed the baseline SMR to increase above 100 because the hypothesis that SMR =100 could be rejected. The Environmental Protection Agency did not report on the fit of their model to the data. On the basis of their meta-analysis, they estimated an ED01 of 47 ng/kg of lipid and 95% lower bound of 30 ng/kg. If the same assumptions as made above (human body, 25% lipid, 50% uptake of TCDD toxic equivalents, and half-life of 7.6 years), this lipid concentration corresponds to a daily intake of 5.9 pg/kg bw per day (95% lower bound, 3.7 pg/kg bw per day).

Steenland et al. (2001) conducted a quantitative risk assessment on the basis of the data of Fingerhut et al. (1991). TCDD concentrations as a function of age were estimated for all 3538 workers, and dose–response analyses were conducted on the basis of the cumulative serum concentration. A significant (p = 0.003) positive dose–response trend was found between the estimated log cumulative TCDD serum concentration and mortality from any cancer. Steenland et al. estimated the additional lifetime risk for cancer due to exposure to TCDD with two models. In one model, the relative risk was assumed to be a linear function of the log cumulative TCDD serum concentration. In the second model, the relative risk was assumed to be a ‘piece-wise’ linear function of the (untransformed) cumulative TCDD serum concentration. The second model is similar to that used by the Committee. It predicted an average additional risk for males and females of 0.00045 at 10 ng/kg over that at 5 ng/kg. As the model is linear, this is equivalent to an ED01 of 5 x 0.01/0.00045 = 111 ng/kg, which, on the basis of the assumptions stated above, corresponds to an intake of 14 pg/kg bw per day. The log-linear model used by Steenland et al. was considered by the Committee to be less plausible than the linear model. For example, the log-linear model predicts that the current 5 ng/kg background serum TCDD concentration is responsible for 44% of all cancer1, and increasing the serum concentration by another 100-fold (from 5 ng/kg to 500 ng/kg) would result in about the same increase in risk as that predicted for the current background of 5 ng/kg. The Committee obtained similar results by applying a similar log-linear model to the data in Table 35.

8.2.3 Short-term effects other than cancer

In adults, most effects of high doses manifest within a few weeks after ingestion of a sufficient dose of dioxins; developmental effects may not manifest for months or years.

Inferences about the dose–response relationship for some of the manifestations of the Yusho syndrome can be drawn from the data of Kuratsune et al. (1972) and Hayabuchi et al. (1979). These authors studied a group of 146 persons who claimed to have ingested contaminated rice-bran oil produced on 1 day (5 February 1968) and sold in 13.5-l cans. They estimated the amount of contaminated oil ingested by the patients and reported the proportions of patients with various symptoms. In the first paper, they describe chloracne, subdiving the cohort into three classes of intake of oil (< 720 ml, 720–1440 ml, > 1440 ml). In the second, they reported that the smallest and largest total intakes had been 220 ml and 3375 ml, respectively. Thus, the classes of intake can be set at 220–720 ml, 720–1440 ml and 1440–3375 ml. The group with low intake comprised 80 persons, 58 of whom (78%) had symptoms of chloracne. The person with the lowest total intake had severe chloracne. All 45 patients with intermediate intake and all 21 with high intake developed chloracne. On the basis of a graphical interpolation in semi-log space, the ED50 for chloracne due to ingestion of oil in the Yusho incident was estimated to be 74–220 ml.

Kuratsune et al. (1972) reported the intakes and outcomes for women who had ingested the oil while pregnant. The smallest intake (300 ml; two women) was associated with minimal effects on the infants, whereas intakes of 700–1400 ml (five women) were associated with symptoms typical of those seen in adults, and some of the infants were recorded as small-for-date. The highest intake (2600 ml) was associated with a stillbirth, and the woman had severe chloracne, as did another woman who had a stillborn infant but whose intake was not estimated.

An ED50 of 610 ml of oil was inferred from these data for severe chloracne (grades III and IV; Wilson, 1987). A mean intake of 688 ml was reported (Hayabuchi et al., 1979). Subjective and clinical symptoms in 46 persons who had undergone physical examinations in 1970 and had responded to a questionnaire were related to a different subdivision of intake: < 499 ml, 500–999 ml and > 1000 ml. Three symptoms were found to show a dose–response relationship: central nervous system symptoms (headache and dizziness), peripheral neuropathy and numbness of the limbs, and abnormalities of the teeth. Using simple log-linear interpolation, the Committee inferred ED50 values for the last two effects of 625 ml and 875 ml, respectively. As more than half the patients with the lowest intake had symptoms of central nervous system involvement, a reliable ED50 for this response could not be estimated. In addition, all these persons also had ocular discharge. ‘Latent’ intakes were reported for a few patients; the term is not clearly defined but appears to mean the amount estimated to have been ingested before medical attention was sought. The lowest reported latent intake was 121 ml and the highest was 1925 ml (Yoshimura & Hayabuchi, 1985).

The intakes of oil were converted to estimates of PCDF intake on the basis of analyses made of the oil at the time of the incident (Morita et al., 1977; Miyata et al., 1985; Tanabe et al., 1989; Yao et al., 2002). The results are shown in Table 38. The toxic equivalents value of the oil was estimated to be 792 ng/g of oil. As rice-bran oil is reported (Masuda, 1992) to have a density of 0.92 kg/l, the oil produced on 5 February 1968 can be estimated to have contained 730 ng toxic equivalents per ml.

Table 38. Concentrations of dioxin-like contaminants in oil produced on 5 February 1968 in the Yusho incident (ng/g of oil)

Congener

Miyata et al. (1977); Morita et al. (1977)

Tanabe et al. (1989)

Yao et al. (2002)

Meana

TEF

Toxic equivalents (ng/g)

1,2,3,7,8-PeCDF

860

290, 760

140

513 ± 351

0.05

25.6 ±17.5

2,3,4,7,8-PeCDF

1180

1300, 1400

680

1 140 ± 320

0.5

570 ± 160

1,2,3,4,7,8-HxCDF

1780

750, 1360

900

1 198 ± 467

0.1

119.8 ± 46.7

+ 2,3,4,6,7,8-HxCDF

 

 

 

 

 

 

3,3΄,4,4΄-TCB

ND

11 000,

11 000

11 333 ± 577

0.0001

1.1 ± 0.1

 

 

12 000

 

 

 

 

3,3΄,4,4΄,5-PeCB

ND

530, 730

980

747 ± 225

0.1

74.7 ± 22.5

3,3΄,4,4΄,5,5΄-HxCB

ND

27

38

45 ± 22

0.01

0.5 ± 0.2

Total toxic equivalents

 

 

 

 

792

 

ND, not determined

a

Mean of four determinations ± one standard deviation for PCDFs; mean of three determinations ± one standard deviation for PCBs.

Many analyses of the oils involved in the Yusho and Yu-Cheng incidents showed the presence of 2,3,7,8-TCDF; however, Chen et al. (1985) reported that this component was in fact the non-toxic 2,3,4,8 congener. Analyses of tissues from patients from the two incidents usually showed the presence of only two of the three toxic PCDFs: 2,3,4,7,8-PeCDF and 1,2,3,4,7,8-HxCDF. The third congener, 1,2,3,7,8-PeCDF, is cleared rapidly from the body, most disappearing from tissues within a matter of months. This introduces uncertainty into the estimates, and to a greater extent for the Yusho than for the Yu-Cheng incident, because ingestion occurred over a longer time in the latter. Because this congener has a low TEF value, its exclusion does not introduce a large error.

These data allowed the Committee to estimate an ED50 value for chloracne and related symptoms (Table 39), assuming that the oil contained toxic equivalents at a concentration of 7400 ng/l of oil and that the mean body mass of the patients was 55 kg. Although the mean body mass of a population is assumed to be 60 kg, the Committee used 55 kg because 18% of the patients were under the age of 11 and the measured mean body mass of seven adult patients was 55 kg (Hayabuchi et al., 1979). When the actual body masses of individuals were given, they were used. ED50 values were estimated for three effects: severe chloracne, 8300 ng/kg bw; peripheral neuropathy, 8500 ng/kg bw; and dental effects, 11 300 ng/kg bw. In addition, the dose of toxic equivalents corresponding to the mean oil intake of the patients, 688 ml, was estimated to be 9.4 ΅g/kg bw. The group therefore had a slight bias towards more severe symptoms.

Table 39. Chloracne and related symptoms seen after ingestion of contaminated rice oil

Effect

Oil intake (ml)

Intake of toxic equivalents (ng/kg bw)a

Reference

‘Severe’ chloracne

610

8 300

Kuratsune et al. (1972); Wilson (1987)

Peripheral neuropathy

625

8 500

Yoshimura & Hayabuchi (1985)

Dental abnormalities

875

11 300

Yoshimura & Hayabuchi (1985)

Central nervous system symptomsb

< 500

< 7000

Yoshimura & Hayabuchi (1985)

Ocular dischargec

< 500

< 7000

Yoshimura & Hayabuchi (1985)

a

Intake estimated on basis of assumption of 0.75 mg toxic equivalents per ml oil and 55 kg mean body mass

b

ED50 estimated to be between 2 and 7 mg/kg bw; 50% of persons with an intake of 0-499 ml complained of or had such symptoms.

c

ED50 estimated to be < 2 mg/kg; ~ 80% of the cohort complained of this sign.

Using a different method, Ryan et al. (1990), estimated the mean intake of the group to be 2400 ng/kg bw; however, this estimate is based on a disposition model with no assumption of extra storage of PCDF in the liver and is thus an underestimate of the dose by a factor of about 5. After adjustment for hepatic storage, the estimate is 10 000 ng/kg bw. All these values are within the bounds of experimental uncertainty, with a mean of 10 000 ng/kg bw. Thus, the Committee assigned an ED50 for severe chloracne or Yusho syndrome of 10 000 ± 2000 ng/kg bw. This is also the approximate ingested dose at which essentially all persons would have been expected to develop chloracne. Some individuals who ingested 6000–7000 ng/kg did not develop chloracne.

Masuda (1992) estimated the LOEL for severe symptoms to be 4000 ng/kg bw. At an oil intake of 500 ml (6700 ng/kg, as toxic equivalents), all members of the group showed symptoms of meibomian gland involvement. The roughly estimated ED50 for mild chloracne, 74–220 ml, is equivalent to 1000–3000 ng/kg as toxic equivalents.

Consistent with these estimates, Coenraads et al. (1999) estimated the concentra-tions of toxic equivalents for seven men with chloracne who had been engaged in pentachlorophenol production to be 1200–22 000 ng/kg of lipid. Table 2 in the paper by Carrier et al. (1995b) indicates that these values correspond to body burdens of about 500–10 000 ng/kg. In the Seveso incident, the estimated body burdens of people with chloracne ranged from 200 to > 40 000 ng/kg bw (Bertazzi et al., 1998a). In both these incidents, some of the exposure was dermal, indicating that the systemic dose was smaller than the tissue dose.

The largest dose received by the Yusho patients studied by Hayabuchi et al. (1979), 3375 ml, is equal to 45 000 ng toxic equivalents per kg bw; the largest dose reported to have been received during the ‘latent period’ (before presentation to a physician), 1934 ml, is equal to ~ 24 000 ng/kg bw. (These two intakes may or may not refer to the same individual.) Geusau et al. (1999) reported blood lipid concentrations of TCDD in two women who were taken to hospital with chloracne of 144 000 and 26 000 ng/kg of lipid, implying ingested doses of > 100 000 and 20 000 ng/kg bw, respectively. The larger dose is somewhat uncertain, as the toxicokinetics models used to infer body burden from serum lipid concentrations has not been tested for lipid concentrations > 20 ΅g/kg. It is assumed in the model that the distribution among compartments does not change markedly at very high doses, which may not be correct.

Hayabuchi et al. (1979) reported the estimated intakes and intake rates for six persons who were believed to have ingested the smallest amounts of contaminated oil and the six who ingested it at the lowest rate. Of these, five were men, three were women and four were children. The results are shown in Table 40. (As two of the 12 children were under 5 years of age, the estimates of their intake were considered insuffuciently accurate.) Four men and one girl had severe symptoms (clinical grades 3 and 4). The intake estimates indicate a LOEL of toxic equivalets for severe chloracne in adults of 3000 ng/kg bw. The intakes of the five patients (three women, one man and one girl) with mild symptoms (grades 0–2) were estimated to have been 4000–7000 ng/kg bw. Ryan et al. (1990) reported that two of 12 victims of the Yu-Cheng poisoning incident were asymptomatic although they were estimated to have had body burdens of 1000–2000 ng/kg bw.

Table 40. Characteristics of eight adults with smallest intakes of contaminated rice oil

Sex

Age (years)

Weight (kg)

–

 

Total intake

Mean intakeb (΅g/kg bw)

Severity of symptoms

ml

΅g/kg bw

ml

΅g/kg bw

Male

56

57

170

2.2

905

11.9

5.2

3

Male

56

58

235

3.1

348

4.5

3.7

3

Female

46

52

353

5.1

523

7.5

6.2

0

Male

44

57

220

2.9

220

2.9

2.9

3

Male

42

58

11

1.6

565

7.3

3.4

3

Male

30

59

309

4.0

309

4.0

4.0

2

Female

23

50

314

4.7

465

7.0

5.7

1

Female

19

51

314

4.6

465

6.8

5.6

0

Female

8

24

146

4.6

195

6.1

5.3

1

Female

6

19

161

6.4

423

17

11.5

4

Male

2

13

199

11.5

360

21

16

2

Female

1

10

~200

~15

–

–

~ 15

0

From Hayabuchi et al. (1979)

a Amount estimated to have been ingested by patients before seeing a physician

b Calculated from geometric mean of latent and total intakes

Numerous reports have been made of persons with estimated body burdens of < 1000 ng/kg bw who did not develop chloracne. For example, Coenraads et al. (1999) reported that nine men without chloracne who had worked in the same production unit as seven men who developed chloracne had a serum concentration of toxic equivalents (pooled sample) of 400–600 ng/kg of lipid, corresponding to about 100 ng/kg bw. In Seveso, people living in zone A and who did not develop chloracne were found to have a serum concentration of TCDD of 1770–10 400 ng/kg of lipid, corresponding to body burdens of 400–6000 ng/kg. Geusau et al. (1999) reported that three co-workers of two women with chloracne had serum concentra-tions of TCDD of 0.09, 0.1 and 0.86 ng/kg of lipid (~ 200 ng/kg bw) but did not develop chloracne.

In sum, dioxins can have mild effects at toxic equivalent doses of 1000–10 000 ng/kg. Larger doses have more severe and more persistent effects; in the Yusho incident, all patients who received such doses and were examined developed chloracne. Men appeared to be more sensitive to the effects of dioxins than women. The largest doses reported were in the range 50 000–100 000 ng/kg bw; when reports were available, such doses were associated with very severe illness.

The data from the Yusho incident imply that children are somewhat less sensitive to the effects of these contaminants than are adults (see Table 40). The data from the Seveso incident, however, suggest that children may be more sensitive (Bertazzi et al., 1998a), although the possibility of dermal exposure complicates interpretation of these data. The data from the Yusho incident also suggest that males are more sensitive to the effects of dioxins than are females, in contrast to rats. Only scattered observations are available from the Seveso incident and elsewhere (e.g. Geusau et al., 1999).

8.2.4 Estimate of lowest effect level for developmental effects from data on humans

Developmental effects in children who ingested contaminated oil in the Yusho incident were reviewed by Yoshimura (1996). Hyperpigmentation, delayed physical growth and nervous system development were found in essentially all children who also had other symptoms of the Yusho syndrome. Although the growth rate deficiencies continued throughout adolescence, the ultimate height of both boys and girls was normal. No symptoms of neurological deficits were observed, but this has not been studied sytematically. In addition, 13 women who were pregnant when they ingested the contaminated Yusho oil had two stillbirths, and nine of the 11 liveborn infants showed signs of intoxication, including stained skin and ocular discharge. Although size at birth was not known for two infants, four were small and showed more severe dermal symptoms. The two ‘unaffected’ infants had some signs of neonatal jaundice but were otherwise normal. One had been exposed only during the third trimester and the other early in the second trimester; their mothers were estimated to have ingested 300 ml of contaminated oil at those times in their pregnancies (~ 4000 ng/kg bw as toxic equivalents), and one had grade 0 and the other grade 2 symptoms. The other women had ingested 700–2600 ml of oil (10 000–35 000 ng/kg bw) during some or all of their pregnancies. One of the two stillbirths was associated with the largest consumption of oil, and the woman had very severe (grade 4) symptoms. No information was available on the oil consumption of the second women who had a stillbirth; the severity of her symptoms was recorded as grade 3.

The development of the children of mothers affected in the Yu-Cheng incident has been studied by Hsu and colleagues (Rogan et al., 1988; Chen et al., 1992; Guo et al.,1993; Yu et al., 1994; Hsu et al., 1995; Lai et al., 2001). Most of the studies were conducted on a cohort of 104 children of women who had ingested contaminated oil and were enrolled in the Yu-Cheng registry, but who had been conceived after their mothers had stopped ingesting the oil. As some of the children had been breastfed, their exposure is considered to have been perinatal. Compari-sons were made to a reference cohort of 104 children, who were closely matched with respect to age, sex, length and weight at birth and socioeconomic status and were drawn from the same geographical area.

These studies confirm and extend the findings from the Yusho incident. In addition, however, deficits in intellectual development that persisted throughout adolescence were reported (Lai et al., 2001). The ultimate physical development of these children has not yet been reported. Additional effects found in perinatally exposed children included effects on the development of fingernails and of the sex organs (e.g. penis length); however, overall sexual development was not delayed. The reduced penile length was statistically significant only for boys born in 1980–82, 3 years after their mothers had ingested contaminated oil.

Some of the effects disappeared with time (hyperpigmentation, neuropathy), while others did not (growth retardation, effects on fingernails, penis length) (Hsu et al., 1995). Although some of these effects are associated only with high doses, consistent with the types of effects seen at effective doses in adults, some may represent developmental effects resulting from prenatal exposures to doses lower than those required to elicit symptoms in the mothers. Rogan et al. (1988) stated that an increased incidence of premature death, both pre- and postnatally, occurred among children of Yu-Cheng as well as Yusho patients, but they gave no data.

Ryan et al. (1994) concluded that the effects were not a result of postnatal exposure, as the responses did not correlate with the children’s body burdens of PCDFs or PCBs. In 1985, when the mean age of the children was 3.6 years, their mean serum concentration of toxic equivalents was 200 ng/kg of lipid, which corresponds to a body burden of 50 ng/kg bw. The authors concluded that the doses received by the mothers would provide a better estimate of the effective body burdens of the affected children, but they did not give the estimates.

The body burdens of the mothers can be estimated from data reported during the time the children were born. Estimates of the blood lipid concentrations of 2,3,4,7,8-PeCDF and 1,2,3,4,7,8-HxCDF of the Yu-Cheng patients can be converted to toxic equivalents from the WHO TEFs and Table 2 of Carrier et al. (1995b), as shown above. On the basis of mean concentrations of 10 000 ng/kg for 2,3,4,7,8-PeCDF and 30 000 ng/kg for 1,2,3,4,7,8-HxCDF for 67 persons in 1980, the mean population body burden of toxic equivalents in that year can be estimated to have been 6000 ng/kg bw. Similarly, on the basis of measurements made on samples taken in 1983, the mean mean population body burden of toxic equivalents in that year can be estimated to have been 1200 ng/kg bw. An approximate lower bound of the body burden in those years can be estimated from the body burden associated with development of chloracne in female victims of the Yusho incident (> 6000 ng/kg bw) and the estimated depuration rates (Carrier et al., 1995b). The minimum body burdens of toxic equivalents in 1980 and 1983 can be estimated to have been 3000 and 370 ng/kg bw, respectively. A rough check on this estimate can be made from the observation that the minimum recorded dose in the Yusho incident was about one-fourth the mean; this suggests that the minima in 1980 and 1983 would have been approximately 1500 ng/kg bw and 300 ng/kg bw, respectively. As severe chloracne is somewhat disabling, women with body burdens at the lower end of this range would probably have been more likely to become pregnant. Thus, the mothers of children in the cohort studied by Guo et al. (1993) can be estimated to have had body burdens of 1500–6000 ng/kg bw in 1979–81 and 300–1200 ng/kg bw in 1982–84.

The reduction in body burdens seen between 1980 and 1983 continued through 1991 (Ryan et al., 1994). As the mean concentration of toxic equivalents decreased between the years of birth of the first and latest age groups of Guo et al. (1993), some dependence of the effects on the year of birth would be expected. Such dependence is observed for penile length but not for other effects. The deficits in sex organ development reported in experimental animals (see section 2.2.5) could be considered similar to the abnormalities in penis length and spermatogenesis reported by Guo et al. (1993). Nevertheless, the estimated body burdens of the mothers in the Yu-Cheng incident (300–6000 ng/kg bw) are 6–120 times those (25–50 ng/kg bw) associated with the effects in rats that were used as the basis of the safety assessment.

Studies of the Seveso incident showed no apparent increase in the incidence of either reproductive effects in women with a median body burden of 0.1 ΅g/kg bw or developmental deficits in infants born to these women (Bertazzi et al., 1998a).

Recent studies of the general population, and in particular groups in The Netherlands, have been cited as providing evidence that pre- or postnatal exposure has effects on neurological and other end-points in children. However, the associations reported in these studies are subtle (and often transient) changes that are within the range of normal variation. The current data support a link with prenatal exposure to total PCBs and not with estimated concentrations of toxic equivalents (Patandin et al., 1999). For example, Patandin et al. (1999) found that raised concentrations of PCBs in maternal plasma (> 3.0 ΅g/L) were associated with lower cognitive scores for all infants and for formula-fed infants, but they found no effect of prenatal or lactational exposure to toxic equivalents on performance. Two end-points, transient, subtle changes in measures of neonatal neurological development and alterations in thyroid hormone concentrations, showed some relationship with toxic equivalents (Pluim et al., 1993; Huisman et al., 1995) in children at the higher end of the exposure spectrum. Furthermore, even the reported associations with prenatal exposure to PCBs were overwhelmed by the beneficial effects of breastfeeding, even though breastfeeding results in much higher doses of PCBs, dioxins and furans for the infant. Thus, these studies do not provide a NOEL or LOEL that can be used in assessing possible tolerable levels of intake of toxic equivalents.

No frank effects, such as birth defects or chromosomal abnormalities, were found in infants of women exposed in Seveso, where the median body burden in zone A was about 110 ng/kg bw on the basis of a median concentration of TCDD of 440 ng/kg of lipid in blood samples taken soon after the accident and an assumption of 25% body fat (Bertazzi et al., 1998a). Exposure in zone A did vary widely, however, and some persons had effects including chloracne, transient increases in liver enzyme activity and transient neurological effects (Bertazzi et al., 1998a). The median peak exposure in zone B has been estimated to have been about 90 ng/kg of lipid, corresponding to a body burden of TCDD of about 22 ng/kg bw. Addition of background concentrations of other congeners that contribute to total toxic equivalents (estimated background body burden in 1976, 6–8 ng/kg bw) would increase the total body burden of toxic equivalents in this group to 28–30 ng/kg bw. No overt effects were observed in this population.

The data on dose–response relationships in the infants of women exposed in Seveso and in the contaminated oil poisoning incidents are shown in Table 41.

Table 41. Dose–response relationships for effects on infants of mothers exposed to dioxins

Population

Findings

Estimated maternal body burden of toxic equivalents (ng/kg bw)

Reference

Seveso, zone B

Few if any detectable effects in adults; no increase in incidence of birth defects

28–30 (median)

Bertazzi et al. (1998a)

Seveso, zone A

Transient liver enzyme and neurological changes; chloracne in some persons; no increase in incidence of birth defects

110 (median)

Bertazzi et al. (1998a)

Yusho and Yu-Cheng incidents

Frank effects on offspring

> 2000

Guo et al. (1995)

8.2.5 Dose–response relationships above the tolerable intake based on data for experimental animals

The dose–response relationship described by Rozman (2000) for fatal anorexia in female rats is linear in semi-logarithmic space (between ~5% and ~ 95% response) and very steep, the entire range of response being encompassed by 20–40 ΅g/kg bw expressed as toxic equivalents. Limited evidence suggests that fatal anorexia in female rhesus monkeys follows a similar function. The steepness of the function suggests a very narrow dispersion of susceptibility for lethality within the populations studied.

Dose–response functions in experimental animals are also available for non-quantal end-points. Enzyme induction, a relatively simple effect known to be related directly to binding to the Ah receptor, has been studied in rats given various doses of TCDD (van den Heuvel et al., 1994). Significant induction of EROD activity was observed in the livers of rats given a single dose of about 0.01 ΅g/kg bw, and the activity increased with dose up to about 1 ΅g/kg bw before flattening out (a sigmoidal-shaped dose–response curve). In contrast, as discussed above, no evidence of increased CYP enzyme activity was found in workers with body burdens > 0.4 ΅g/kg bw (Halperin et al., 1995). This difference in responsiveness to TCDD of an end-point as basic as enzyme induction supports the conclusion that humans may be less responsive than laboratory rodents for a variety of end-points.

8.2.6 ‘Benchmark’ dose modelling

Selected experimental data sets were analysed by ‘benchmark’ dose modelling to obtain a quantitative assessment of the body burdens associated with selected effects in laboratory animals. The ED10 was calculated with ‘Benchmark Dose’ software. Most of the responses considered were reported as continuous data. For these, an abnormal response was defined by the estimated 5th percentile of response in unexposed animals, so that the estimated 5% of the most severe effects seen among unexposed animals was considered to be adverse.

The data sets modelled were selected from those compiled by the Environmental Protection Agency in the USA during its re-assessment of dioxins and do not necessarily represent the full range of toxicological data available. The data sets were selected to represent the end-points of interest on the basis of a review of all the toxicological data. For instance, data from the studies of Gray et al. (1997a,b) and Mably et al. (1992a,b) on developmental effects in male rat offspring were included, but some data sets of interest were not used, such as data from Gehrs & Smialowicz (1999) on end-points in the immune system of rat offspring exposed in utero and during lactation.

About 180 data sets were analysed with the power and Hill dose–response models. The power model is a limiting case of the Hill model and was used unless the Hill model provided a significantly better fit. Similarly, the standard deviation was assumed to be independent of dose, unless a contrary assumption give a significantly better fit (higher likelihood). The Hill model includes four disposable parameters, which makes it quite flexible. As a result, however, it tended to gave unreasonable (extremely low) values for some data sets. An example can be generated by applying the Hill model to the data of Theobald & Peterson (1997) on mouse epididymal weights. The epidydimal weights did not vary by dose at single doses of 15 000, 30 000 and 60 000 ng/kg bw, nor were they statistically significantly different from those of controls. Fitting of the Hill model resulted in an ED10 value of < 1 ng/kg bw for this end-point, which is more than four orders of magnitude lower than the lowest dose tested. This result is clearly an artifact of the modelling and does not provide biologically meaningful insight. When the power law was used to model this data set, the estimated ED10 value was about 30 000 ng/kg bw, which is a more biologically meaningful estimate that is more consistent with the toxicological data. Because of this feature of the Hill model, results obtained with this model were considered to be less reliable than those obtained with the power model, which has only three parameters. Consequently, the results are summarized in two ways: with both the Hill and power models as described above, and with only the power model (even if the Hill model gave a better fit).

Doses were converted into steady-state body burdens for the 10% response for comparison among datasets. The conversion formula used was:

Estimated body burden for 10% response (ng/kg bw) = ED10 (ng/kg bw per day) Χ half-life (days)/ ln(2) Χ F,

where ED10 is the calculated extra risk for the effect, the half-life is specific to species (see Table 42) and F is the fraction of TCDD absorbed (gavage = 1; feed = 0.5).

Table 42. Results of modelling data sets from experiments in animals to estimate effective doses associated with a 0.1 increase in the probability of an abnormal response (ED10 )

Species

No. of data sets

ED10 with fits to Hill or power model (ng/kg bw)

ED10 with fits to power model only(ng/kg bw)

Median (range)

10th percentile

Median (range)

10th percentile

Developmental effects in rodent offspring (single doses)

Rat

54

584 (0.11–22 200)

21

626 (208–38 300)

317

Mouse

42

35 700 (0.79–190 300)

228

47 122 (216–190 000)

6 140

Effects on immune system (single doses)

Mouse

27

15 000 (53–88 800)

130

20 600 (474–88 800)

923

Effects on immune system (multiple doses)

Mouse

7

30 900 (457–80 500)

1 120

45 900 (15 900–80 500)

18 800

Effects on body and organ weights (single doses)

Rat

4

49 300 (3457–100 800)

4 762

61 200 (7808–101 000)

15 000

Mouse

6

74 600 (28 522–127 000)

42 800

74 600 (28 500–127 000)

22 900

Effects on body and organ weights (multiple doses)

Rat

22

4420 (63–57 000)

116

6450 (917–57 000)

2 870

Effects on concentrations of thyroid hormones (multiple doses)

Rat

4

3890 (245–5520)

1 078

3880 (1487–5520)

1 950

The modelled data sets were derived exclusively from published studies in mice and rats. Data on strains studied for their particular sensitivity or resistance to TCDD were not used. The end-points modelled were changes in development, the immune system, thyroid hormone concentrations and body and organ weights. Most of the data sets were from experiments in which a single dose was administered; however, some data sets from studies with multiple dosing were also modelled.

Table 42 summarizes the results of the benchmark dose modelling with both approaches, and Figures 30–32 present the ED10 values obtained with the two models efforts in box-and-whisker plots. The values are presented as body burden (in ng/kg bw) as estimated from the dosing regimens used in the studies. For studies with single doses, the value presented is the peak body burden attained after dosing. For studies with multiple doses, the values are estimates of the average body burden. Table 42 presents the median and range of estimated ED10 values by category of effect and species. It also presents the 10th percentile of the estimated ED10 values for each category and species. The 10th percentile is reported because it represents a reasonably stable estimate of the lower end of the estimated values in each category.

FIGURE 30

Upper panel, power of Hill model; lower panel, power model only. The boundary of the box closest to 0 represents the 25th percentile, and the boundary furthest from 0 represents the 75th percentile. The whiskers represent the 90th and 10th percentiles. Values above the 90th percentile and below the 10th percentile are shown as individual points.

Figure 30. ED10 values for immune effects

FIGURE 31

Upper panel, power of Hill model; lower panel, power model only. The boundary of the box closest to 0 represents the 25th percentile, and the boundary furthest from 0 represents the 75th percentile. The whiskers represent the 90th and 10th percentiles. Values above the 90th percentile and below the 10th percentile are shown as individual points.

Figure 31. ED10 values for developmental effects

FIGURE 32

Upper panel, power of Hill model; lower panel, power model only. The boundary of the box closest to 0 represents the 25th percentile, and the boundary furthest from 0 represents the 75th percentile. The whiskers represent the 90th and 10th percentiles. Values above the 90th percentile and below the 10th percentile are shown as individual points.

Figure 32. ED10 body burdens for effects on body and organ weights

In all cases, the results of the modelling neglected the body burdens of toxic equivalents present in control animals. As demonstrated by van den Heuvel et al. (1994) for rats and by DeVito et al. (1995) for mice, control laboratory rodents have body burdens of toxic equivalents of 2–6 ng/kg bw, depending on the age of the animal. Because of these background body burdens, benchmark doses should be interpreted as incremental doses that add to the existing body burdens.

In general, the lowest ED10 values were derived for developmental effects in rats in the studies of Gray et al. (1997a,b) and Mably et al. (1992a,b). With the exception of these values, the ED10 values estimated with the Hill model were generally 100 ng/kg bw or more, and all the median values for the various categories of effect were > 500 ng/kg bw. In contrast, modelling only with the power model resulted in somewhat higher estimates of the ED10, with all median values for various categories > 600 ng/kg bw and, more strikingly, the lowest values > 200 ng/kg bw.

As the power and Hill models are either linear or sublinear, a conservative (low) estimate of the ED01 can be made by dividing the ED10 by 10. The estimated ED01 values were > 20 ng/kg bw. By comparison, the human body burden corresponding to the recommended EHMI is 3–7 ng/kg bw.

The benchmark analysis provides a useful perspective on the NOELs identified by the Committee in specific studies and on the recommended EHMI. However, the exercise also demonstrates that the results of such modelling are highly model-dependent. Therefore, the results of any modelling effort should be re-examined in light of the toxicological data. These issues should be studied further before the results of such modelling can be applied reliably in safety assessment.

9. COMMENTS

9.1 Toxicokinetics

9.1.1 Absorption and biotransformation

Coplanar compounds in dietary fat pass easily from the gut into the blood. Indeed, experiments in humans and laboratory animals given an oral dose of TCDD showed 50–90% absorption. This figure is comparable with the near-complete absorption of PCDDs, PCDFs and PCBs by nursing infants from their mothers’ milk.

After absorption from the gut, TCDD enters the lymph in the form of chylomicrons and is cleared from the blood within 1 h, to appear mainly (74–81% of an administered dose) in the liver and adipose tissue. After clearance from the blood, coplanar compounds remain mainly in serum lipoproteins (very low density, low density and high density), and some are bound to serum proteins.

The Committee used the results of a study in which [3H]TCDD was given to pregnant Long-Evans rats by gavage at a dose of 50, 200, 800 or 1000 ng/kg of body weight on day 15 of gestation, and the radiolabel was measured in tissues 1 day after treatment. The average maternal body burdens (with the percentage of the dose in the four treatment groups) were 31 (60%), 97 (48%), 520 (65%), and 580 (59%) ng/kg of body weight, respectively. On the basis of this study, the Committee used a value of 60% for the percentage of TCDD retained in pregnant rats 1 day after administration of a single dose by gavage on day 15 of gestation.

The distribution of PCDDs and PCDFs between the blood and organs is governed by lipid partitioning and binding to plasma proteins. The concentrations of PCDDs and PCDFs in blood and adipose tissue are closely correlated. TCDD is distributed between blood and adipose tissue by lipid partitioning, whereas the distribution of HxCDDs, HxCDFs, OCDDs and OCDFs is also governed by binding to plasma proteins.

Binding to plasma proteins plays an important role in the uptake of coplanar compounds from the blood in the liver, even for lower chlorinated congeners. When rodents are exposed to increasing doses of TCDD, it is preferentially sequestered in the liver. After entering liver cells, TCDD either dissolves in the lipid fraction or binds to the Ah receptor or CYP proteins, probably microsomal CYP 1A2. As the amounts of CYP 1A and CYP 1B proteins in cells are regulated by formation of the TCDD–Ah receptor complex, exposure to increasing amounts of TCDD results in increased formation of this complex, which leads to increased production of CYP 1A and CYP 1B mRNA and proteins (enzyme induction), and accumulation of TCDD by increased binding to the induced CYP proteins. Similar sequestration has been observed with higher chlorinated PCDDs and PCDFs and with coplanar PCBs.

The hepatic sequestration of coplanar compounds markedly affects their distribution in the body. For example, whereas the liver usually contributes 10% and the adipose tissue 60% of the body burden of TCDD in mice, these fractions may increase to 67% in liver and decrease to 23% in adipose tissue in mice in which hepatic CYP proteins have been fully induced. Similar results were found in rats, clearly indicating the non-linear character of the kinetics of TCDD at concentrations that induce hepatic CYP proteins.

As in rodents, preferential sequestration of PCDDs and PCDFs in the liver rather than in adipose tissue has been observed in humans exposed to background concentrations of these compounds. Although Ah receptor-dependent CYP induction has been observed in human liver cells in vitro after exposure to TCDD, it occurred at concentrations that were several orders of magnitude higher than those observed in human blood. It is therefore likely that the sequestration is due to binding to constitutive CYP proteins.

9.1.2 Metabolism and excretion

In laboratory animals, PCDDs and PCDFs are excreted almost exclusively in the bile, excretion in the urine being a minor route. Whereas the parent compound is found primarily in the organs of rodents, only metabolites of PCDDs and PCDFs occur in bile, indicating hepatic metabolism, including hydroxylation and conjugation, of these compounds. Similar reactions were found in vitro when recombinant human CYP 1A1 was incubated with TCDD. Faecal excretion of unmetabolized PCDDs and PCDFs is also an important route of elimination in humans.

In rodents, the half-life of TCDD ranges from 8–24 days in mice to 16–28 days in rats. Humans eliminate PCDDs and PCDFs more slowly, the estimated mean half-life of TCDD ranging from 5.5 to 11 years. The half-lives of other PCDD congeners and of PCDFs and coplanar PCBs vary widely. These differences in the half-lives of different congeners are reflected in their TEFs (see Table 1).

9.1.3 Relationship between human intake and doses used in studies in laboratory animals

The biochemical and toxicological effects of PCDDs, PCDFs and coplanar PCBs are directly related to their concentrations in tissues, and not to the daily dose. The most appropriate measure of dose would therefore be the concentration at the target tissue; however, this is seldom known. The body burden, which is strongly correlated with the concentrations in tissue and serum, integrates the differences in half-lives between species. Thus, rodents require appreciably higher daily doses (100–200-fold) to achieve a body burden at steady state that is equivalent to that recorded in humans exposed to background concentrations. Toxicokinetically, estimates of body burden are therefore more appropriate measures of dose for interspecies comparisons than is the daily dose.

The long half-lives of PCDDs, PCDFs and coplanar PCBs have several implications for the period of intake that is relevant to the assessment. First, the concentration of toxic equivalents in the body (or the internal toxic equivalents to which a target organ is exposed) will increase over time as more of the compounds are ingested. Second, after cessation of exposure, the body’s concentration of stored toxic equivalents (and the exposure of internal organs) will decline slowly, only half of the accumulated toxic equivalents disappearing over about 7 years, resulting in a pseudo-steady state only after decades. Third, because of this long-term storage in the body and the consequent daily exposure to the body’s stored toxic equivalents, intake on a particular day will have a small or even negligible effect on the overall body burden. For example, in the unlikely event of food contamination that leads to an intake 100 times the amount present in a typical meal, the body burden of the adult eating that meal would increase by < 3%. The rest of the body burden would be made up of the PCDDs, PCDFs and coplanar PCBs consumed in many thousands of meals over the previous decade or more.

Therefore, the Committee concluded that the appropriate period for evaluating the mean intake of these compounds is 1 month.

In order to transform an animal body burden into an EHMI that on a long-term basis would result in a similar body burden (at steady state), simple, classical toxicokinetic calculations can be used. The elimination of low doses of PCDDs was considered to follow first-order kinetics and to be independent of the body burden or dose. The Committee calculated the total body burden at steady-state using the following equation:

EQUATION 1

where f is the fraction of dose absorbed from food (assumed to be 50% in humans) and the estimated half-life of TCDD is 2774 days (7.6 years). For compounds that follow first-order kinetics, four to five half-lives will be required to approach steady state. For TCDD, this would be equivalent to more than 30 years.

This model is based on the assumption that PCDDs are distributed in only one compartment: the whole body. Although most of the body burden of PCDD is distributed in the lipid stores, at higher doses the liver also sequesters these compounds in both humans and animals. Predictions of body burden after intake of high doses that are based on lipid concentrations may therefore be underestimates of the total body burden (and the intake leading to that body burden), because of hepatic sequestration. Use of physiologically based pharmacokinetic models may be more appropriate under these circumstances. In order to transform the body burdens resulting from intake of the low concentrations to which the general population is exposed and from the low doses used in the pivotal toxicological studies into estimated human daily intake, the Committee considered use of a less complicated, classical pharmacokinetic model to be appropriate.

9.1.4 Determinants of dose received by fetuses in studies of developmental toxicity

The time of dosing in several of the studies considered by the Committee, day 15 of gestation, marks the onset of the sensitive phase of sexual differentiation in rats and represents a critical time of fetal exposure. The determinant of the reproductive effects is the fetal concentration on days 15–16 of gestation, which in turn is determined by the maternal serum concentration. The latter concentration differs with a bolus dose (as in these studies) and with repeated doses providing the same total intake. As the serum concentration of TCDD after a bolus dose rises before distribution to the tissue compartments, the serum concentration is likely to be higher than that after long-term intake of a lower concentration.

The difference in the fetal body burden after a single bolus dose and after repeated administration of low doses resulting in a similar maternal body burden was addressed in a study in Long-Evans rats treated on day 16 of gestation (Hurst et al., 2000a,b). The rats were given [3H]TCDD at 1, 10 or 30 ng/kg of body weight per day by gavage in corn oil, on 5 days per week for 13 weeks. They were then mated, and dosing was continued daily throughout gestation. The regimen produced a steady-state concentration of TCDD in the dams. The average maternal and fetal body burdens on day 16 of gestation after this treatment and after administration of a single dose of TCDD by gavage on day 15 of gestation are shown in Table 43.

Table 43. Average maternal and fetal body burdens after a single dose and after administration of repeated doses of TCDD to pregnant Long-Evans rats

Dose (ng/kg bw per day)

Body burden on day 16 of gestation (ng/kg bw per day)

Maternal

Fetal

Single dose

50

30

5.3

200

97

13

800

520

39

1000

520

56

Repeated dosesa

0.71

20

1.4

7.1

120

7.5

21

300

15

From Hurst et al. (2000a,b)

a

Daily dose, adjusted for continuous administration from 5 to 7 days per week

As expected, a single dose on day 15 of gestation by gavage resulted in considerably higher fetal concentrations on day 16 than short-term administration of low daily doses leading to maternal steady-state body burdens of similar magnitude.

Using the data in Table 43, the Committee conducted least-squares linear fits of dose versus maternal and fetal body burdens. Since radiolabelled TCDD was used in both studies, a zero intercept was assumed for the fitted line. None of these fits showed what appeared to be any significant deviation from linearity. These data indicate that the ratio of fetal:maternal body burden resulting from a bolus dose would be 1.7 times that from multiple doses providing the same total dose. Kinetic data indicate that a linear dose–response relationship would be expected at the doses used in these studies. The fetal and maternal body burdens in both data sets were also fitted to power equations, which provided a better fit of the data obtained at the lower end of the range of single doses. The factor used to convert maternal body burden after single doses to a corresponding steady-state body burden with the power equations was 2.6.

9.2 Toxicological studies

9.2.1 Acute toxicity

In laboratory animals, the acute toxicity of TCDD and related PCDDs and PCDFs substituted in at least the C-2, C-3, C-7 and C-8 positions varies widely between and among species. For example, the median lethal dose in guinea-pigs treated orally was 0.6 ΅g/kg bw, while that in hamsters was > 5000 ΅g/kg bw. Explanations for this variation include differences in Ah receptor functionality (size, transformation and binding of the PCDD response element), toxicokinetics (metabolic capacity and tissue distribution) and body fat content. While data on acute toxicity were available for various commercial PCB mixtures (median lethal doses usually > 100 mg/kg bw), the data on individual coplanar PCB congeners in mammals were limited.

One of the commoner symptoms associated with lethality induced by PCDDs is a generalized delayed wasting syndrome characterized by inhibition of gluconeo-genesis, reduced feed intake and loss of body weight. Other toxic effects observed after a single exposure to PCDDs include haemorrhages in a number of organs, thymic atrophy, reduced bone-marrow cellularity and loss of body fat and lean muscle mass, although some differences in the frequency of these effects was seen among species.

9.2.2 Carcinogenicity

TCDD and other PCDDs induced tumours at multiple sites in laboratory animal species of each sex. In a series of assays in vivo and in vitro, TCDD promoted the growth of transformed cells (e.g. rat tracheal epithelium cells treated with N-methyl-N΄-nitro-N-nitrosoguanidine), consistent with observations of cancer promotion in whole animals in vivo. In a long-term study of carcinogenicity with TCDD in rats, the LOEL for hepatic adenomas in females was 10 ng/kg bw per day, and the NOEL was 1 ng/kg bw per day. Several studies have shown that TCDD promotes tumours in laboratory animals, in particular liver tumours. Several other PCDDs, PCDFs and non-ortho- and mono-ortho-PCBs also promoted liver tumours. In a long-term study in rats in which the incidence of liver tumours was increased over that in controls, the LOEL of 10 ng/kg bw per day corresponded to a steady-state body burden of 290 ng/kg bw. In order for humans to attain a similar steady-state body burden, they would have to have a daily intake of 150 pg/kg bw (see Eq. 1 in 9.1.3).

9.2.3 Genotoxicity

The results of several short-term assays for genotoxicity with TCDD, covering various end-points, were negative. Furthermore, TCDD did not bind covalently to DNA from the liver of mice. The Committee concluded that TCDD does not initiate carcinogenesis.

9.2.4 Developmental toxicity

A number of biochemical changes, including enzyme induction, altered expression of growth factors and enhanced oxidative stress, have been noted in laboratory animals with body burdens of TCDD within a lower range of 3–10 ng/kg bw. The Committee considered these biochemical effects to be early markers of exposure to PCDDs, PCDFs and coplanar PCBs, or events induced by these compounds in animals and in humans that may or may not result in adverse effects at higher body burdens.

The Committee reviewed the relevant studies (Mably et al., 1992a,b,c; Rier et al., 1993; Gray et al., 1997a,b; Gehrs et al., 1997) considered by the WHO consultation held in 1998 (van Leeuwen & Younes, 2000), as well as three recent studies (Faqi et al., 1998; Gehrs & Smialowicz, 1999; Ohsako et al., 2001). The Committee noted that the most sensitive adverse effects reported were on development in the male offspring of rats and immunological deficits in rats after prenatal exposure to TCDD (see Table 44).

Table 44. Studies in which the lowest NOELs and LOELs were identified for the most sensitive adverse effects of TCDD on developmental end-points in ratsa

Dosing regimen (references)

Strain

End-point

NOEL body burden (ng/kg bw)

LOEL body burden (ng/kg bw)

Single bolus by gavage on day 14 of gestation (Gehrs et al., 1997; Gehrs & Smialowicz, 1999)

Fischer 344

Immune suppression in offspring

 

50

Single bolus by gavage on day 15 of gestation (Ohsako et al., 2001)

Holtzman

Reduced entral prostate weight; decreased ano- genital distance in male offspring

13

51

Single bolus by gavage on day 15 of gestation (Mably et al., 1992c)

Holtzman

Decreased sperm count in offspring

 

28

Single bolus by gavage

Long-Evans

Accelerated eye opening and decreased sperm count in offspring on day 15 of gestation (Gray et al., 1997a)

 

28

Loading and maintenance doses by subcutaneous injection(Faqi et al., 1998)

Wistar

Decreased sperm production and altered sexual behaviour in male offspring

 

25

a Body burdens estimated from a linear fit to the data in Table 43

The WHO consultation identified a study in which endometriosis was found after long-term administration of TCDD to rhesus monkeys. The Committee stressed that the findings in this study should be interpreted with caution, as the daily intake was not adequately reported. In addition, analyses conducted 13 years after the end of exposure showed high concentrations of coplanar PCBs in the blood of the monkeys with endometriosis, possibly from an an unknown source. The Committee also noted that the LOELs in some of the pivotal studies in rats (Table 44) would result in EHMIs that were similar to or lower than that derived from the LOEL for endometriosis in monkeys.

In a recent study (Ohsako et al., 2001), pregnant Holtzman rats were given a single oral dose of TCDD at 0–800 ng/kg bw on day 15 of gestation, and the male offspring were examined on days 49 and 120 after birth. No changes were seen in testicular or epididymal weight nor in daily sperm production or sperm reserve at any dose. However, the weight of the urogenital complex, including the ventral prostate, was significantly reduced at doses of 200 and 800 ng/kg bw in rats killed on day 120. Moreover, the anogenital distance of male rats receiving doses > 50 ng/kg bw and killed on day 20 was significantly decreased. The Committee noted that administration of TCDD at any dose resulted in a dose-dependent increase in 5alpha-reductase type 2 mRNA and a decrease in androgen receptor mRNA in the ventral prostate of rats killed at day 49 but not in those killed at day 120, with no adverse sequelae at the lowest dose of 12.5 ng/kg bw. On the basis of 60% absorption and an assumption of a linear relationship for the data in Table 43, the equivalent maternal body burden after multiple doses at this NOEL would be 13 ng/kg bw. Fitting the data in Table 43 into the power equation, the Committee estimated the body burden NOEL to be 19 ng/kg bw. The LOEL of 50 ng/kg bw per day corresponds to an equivalent body burden of 51 ng/kg bw with the linear model and 76 ng/kg bw with the power model.

The lowest LOEL reported for the reproductive system of male offspring was found in an experiment with Wistar rats (Faqi et al., 1998). In this study, the dams were treated subcutaneously before mating and throughout mating, pregnancy and lactation. They received an initial loading dose of [14C]TCDD at 25, 60 or 300 ng/kg bw 2 weeks before mating, and then a weekly maintenance dose of TCDD at 5, 12 or 60 ng/kg bw. The size of the maintenance doses was determined on the basis of a reported elimination half-life for TCDD of 3 weeks in adult rats. The effects on male reproductive end-points were studied on days 70 and 170 after birth. The number of sperm per cauda epididymis at puberty and in adulthood was lower in the offspring of all treated dams than in those of controls. Daily sperm production was permanently lower in offspring of treated dams than in those of controls, as was the sperm transit rate, thus increasing the time required by the sperm to pass through the cauda epididymis. Moreover, the offspring of the treated groups showed increased numbers of abnormal sperm when investigated in adulthood. The latency periods to mounting and intromission were significantly greater in offspring of dams at the lowest and highest doses, but not of those at the intermediate dose, than in offspring of controls. The Committee noted the lack of clear dose–response relationships for most of these effects in the treated groups. In the male offspring of dams at the highest dose, the concentration of serum testosterone was decreased in adulthood, and permanent changes found in the testicular tubuli included pyknotic nuclei and the presence of cell debris in the lumen. The fertility of the male offspring was not affected in any of the treated groups.

In computing the long-term dose required to produce the fetal concentration found in the group given the initial loading dose of 25 ng/kg bw, the Committee noted that the dose would have been reduced to 20 ng/kg bw before the maintenance dose of 5 ng/kg bw given on day 14. On the basis of the linear fit to the data in Table 43, the fetal body burden resulting from the maternal body burden of 20 ng/kg bw would be 1.04 ng/kg bw. The maintenance dose of 5 ng/kg bw administered on day 14 of gestation would make an additional contribution to the fetal body burden of 0.27 ng/kg bw, resulting in a total fetal body burden of 1.31 ng/kg bw. On the basis of a linear fit to the data in Table 43, a maternal body burden of TCDD of 25 ng/kg bw at steady state would be required to produce this fetal body burden.

The studies summarized in Table 44 provide evidence that adverse effects on the reproductive system are induced in male offspring of pregnant rats given TCDD. The studies show reductions in daily sperm production, in the number of sperm in the cauda epididymides and in epididymal weight as well as accelerated eye opening, a reduction in anogenital distance and feminized sexual behaviour in male offspring associated with maternal steady-state body burdens of TCDD of > 25 ng/kg bw. Reductions in the weights of the testes and the size of the sex accessory glands, such as the ventral prostate, in male offspring, development of external malformations of the genitalia in female offspring and reduced fertility in males and females required higher maternal body burdens.

The Committee noted that the most sensitive end-points differed between studies, perhaps reflecting strain differences in sensitivity and even minor differences in the experimental conditions, e.g. the diet. The Committee also noted that, in one study, administration of a single dose of TCDD at 12.5 ng/kg bw to dams by gavage decreased the amount of androgen receptor mRNA in the ventral prostate of offspring at puberty on day 49 after birth, indicating reduced androgenic responsiveness. However, none of the other above-mentioned adverse effects was seen in male offspring at this dose, which corresponds to an estimated maternal steady-state body burden of TCDD of approximately 19 ng/kg bw (Table 44). The Committee considered the effect on androgenic responsiveness to be an early marker of exposure to TCDD, like enzyme induction, altered expression of growth factors and enhanced oxidative stress, or an event that may or may not result in adverse effects in animals at higher body burdens.

9.2.5 Observations in humans

(a) Effects other than cancer

In two episodes of food poisoning in China (Province of Taiwan) and Japan, in which infants were exposed in utero to heat-degraded PCBs, a variety of adverse physical developmental abnormalities was observed, including decreased penis length and alterations of spermatozoa; neurodevelopmental abnormalities were also seen. The affected children in Taiwan were born to mothers with estimated body burdens of toxic equivalents of PCBs of 2–3 ΅g/kg bw.

Environmental or background exposure of infants in Germany, The Netherlands and the USA was evaluated in several studies; for example, the mean concentration of toxic equivalents in human milk was 60 pg/g of lipid (range 25–155 pg/g) in a study in Groningen and Rotterdam, The Netherlands. Low birth weight, detriments in neurological development and alterations in thyroid hormones, the distribution of lymphocyte subpopulations and the frequency of infections and respiratory symptoms were observed. The observed neurodevelopmental deficits were subtle and the prevalence was within the normal range; their potential consequences for future intellectual function are unknown. The associations observed were considered to be due to prenatal exposure rather than to postnatal intake (from milk). In one study of breastfed and bottle-fed infants, the intake of PCDDs and PCBs was inversely related to performance in neurobehavioural tests, breastfed infants having better scores than bottle-fed infants. These studies of low exposure related primarily to PCBs, and fewer data were available on the effects of PCDDs and PCDFs.

In adults, most of the effects other than cancer observed after exposure to PCDDs, PCDFs and coplanar PCBs, such as chloracne, appeared only at doses several orders of magnitude greater than those generally received from background contamination of foods. In Seveso, Italy, more female children than expected were born to fathers who had serum TCDD concentrations > 80 pg/g of lipid (16–20 ng/kg bw) at the time of conception.

(b) Carcinogenicity

A working group convened by IARC (1997) classified TCDD as a human carcinogen (Group 1). Other PCDDs and PCDFs were considered not to be classifiable as to their carcinogenicity to humans (Group 3).

The most informative studies for evaluating the carcinogenicity of TCDD are four cohort studies of herbicide producers (two in Germany and one each in The Netherlands and the USA) and one cohort study of residents of a contaminated area in Seveso, Italy. A multi-country cohort study from IARC included three of these four cohorts, other industrial cohorts, many of which had not been reported in separate publications, and a cohort of professional herbicide applicators.

In most of the epidemiological studies considered, exposure had been primarily to TCDD, with some exposure to mixtures of other PCDDs, as contaminants of phenoxy herbicides and chlorophenols. The studies involved persons with the highest recorded exposure to TCDD, the estimated geometric mean blood lipid concentrations after the last exposure ranging from 1100 to 2300 pg/g of lipid in the industrial cohorts; lower average concentrations were found in the population exposed in Seveso.

Low excess risks of the order of 40% were found for all neoplasms combined in all the studies of industrial cohorts in which the exposure assessment was adequate. The risks for cancers at specific sites were increased in some of the studies, but the results were not consistent between studies, and no single cancer site seemed to predominate. The results of tests for trends for increasing excess risks for all neoplasms with increasing intensity of exposure were statistically significant. Increasing risks for all neoplasms with time since first exposure were observed in those studies in which latency was evaluated. The follow-up of the Seveso cohort has so far been shorter than that of the industrial cohorts; however, the rate of death from all cancers has not been found to differ significantly from that expected in the general population. Excess risks were seen for cancers at some specific sites among persons in the most heavily contaminated zones at the time of the accident, but there were few cases.

In these well-conducted cohort studies, the intensity of exposure could be ascertained with precision because of the long biological half-life of TCDD in human tissues, and the relative risks increased significantly with increasing exposure. Although the excess cancer risk at the highest exposure was statistically significant, these results must be evaluated with caution, as the overall risks are not high and the strongest evidence is for industrial populations whose exposure was two to three orders of magnitude greater than that of the general population, who also had heavy exposure to other chemicals; furthermore, lifestyle factors such as smoking were not evaluated. There are few precedents of carcinogens that increase the risk for cancer at all sites combined, with no excess risk for any specific tumour predominating.

A ‘benchmark dose’ was calculated from the effective dose estimated to result in a 1% increase in cancer mortality (ED01), on the basis of a meta-analysis of data for three industrial cohorts with well-documented exposure and comparison with the doses required for effects other than cancer. A statistically significant linear trend in risk with intensity of exposure was observed, which persisted even after exclusion of the groups with the greatest exposure. Within the range of reasonable assumptions, the ED01 differed quite widely and depended strongly on the assumptions made. Furthermore, a number of uncertainties would influence the predicted ED01, including the exact exposure of the occupational cohorts and, to a lesser extent, the potential confounding effects of factors not considered in the studies.

9.3 Sampling and analytical methods

As no specific guidelines have been drawn up for sampling foods to be analysed for their PCDD, PCDF and coplanar PCB content, the basic guidelines for sampling for organic contaminants or pesticides should be used. The objective is to obtain a representative, homogeneous laboratory sample without introducing secondary contamination. Although PCDDs, PCDFs and coplanar PCBs are chemically stable, the samples should be stored and transported in such a way that they do not deteriorate. PCDDs, PCDFs and coplanar PCBs are usually found as complex mixtures of varying composition in different matrices. Their identification and quantification require a highly sophisticated method of analysis in order to separate the toxic congeners listed in Table 1 from the more prevalent, less toxic congeners. Usually, PCDDs, PCDFs and coplanar PCBs are determined by capillary gas chromatography with mass spectrometry.

No official method exists for the determination of these compounds in food. Reliable results have been obtained in the absence of official methods when the method used has been shown to be suitable and to fulfil analytical quality criteria developed in other fields of residue analyses. The methods used to determine PCDDs and PCDFs in food must provide sufficient information to allow calculation of the results as toxic equivalents, at concentrations of 0.1–1 pg/g of fat in milk, meat and eggs, 10 pg/g of fat in fish or > 100 pg/g of fat in cases of heavier contamination, and 0.1–0.5 pg/g of dry matter in food of vegetable origin. The patterns of congeners can vary between regions and foods.

When the method used is of insufficient sensitivity, the concentrations of PCDDs, PCDFs and coplanar PCBs in many foods may be near or below the limit of quantification. The method used to derive the concentrations of undetected congeners (the imputation method) can therefore have a variable effect on the summary toxic equivalent value for a food sample. In the most commonly used imputation methods, the contribution of each undetected congener to the toxic equivalent is considered to be either 0 (‘lower-bound concentrations’), the limit of detection or limit of quantification (‘upper-bound concentrations’) or half the limit of detection or limit of determination. In methods with insufficient sensitivity, the lower- and upper-bound concentrations can differ by a factor of 10–100 or even more. If the sensitivity is appropriate, the differences between lower- and upper-bound concentrations are negligible. Therefore, low estimates of PCDDs, PCDFs and coplanar PCBs in a sample may represent truly low concentrations or be the result of use of zero as the value for undetected congeners in a food sample. Conversely, high estimates may be the result either of actual contamination of the food or of use of the upper-bound concept with insufficient sensitivity.

Application of upper-bound or lower-bound concentrations leads to over- and underestimates of intake, respectively. Therefore, the Committee recommended that laboratories report their results as lower-bound, upper-bound and half-detection limits, in addition to values for individual congeners, thus providing all the necessary information for interpreting the results for specific requirements. Experts who are summarizing results based on toxic equivalents should indicate the way in which the toxic equivalents were calculated.

For analysis of food samples with normal background contamination with PCDDs, PCDFs or PCBs, gas chromatography with high-resolution mass spectrometry has been validated in collaborative studies and has been shown to provide the required sensitivity and specificity. Bioanalytical assays have been developed for rapid screening of sediments, soil, fly ash and various foods, but only the chemical-activated luciferase gene expression (CALUX) assay has been used for food, and validation of this assay has begun. While gas chromatography with mass spectrometry is the most powerful method for identifying and quantifying congeners and for recognizing congener-specific patterns, it does not allow direct measurement in a matrix of all congeners present that act through the Ah receptor pathway. The CALUX assay provides an indication of the toxic equivalents present in a certain matrix, including interactive (synergistic or antagonistic) effects; however, it cannot provide information on the pattern of congeners.

The Committee recognized that the available analytical data on PCDDs, PCDFs and coplanar PCBs are limited by the lack of generally accepted criteria for intra- and inter-laboratory validation. Mutual acceptance of analytical methods would be facilitated by international collaborative studies and proficiency testing programmes. For reliable analysis of concentrations in the range of normal background contamination, laboratories must use sufficiently sensitive methods. General statistical parameters that have been established in other fields of residue analysis could be used. The requirements for acceptable analytical methods clearly need to be harmonized, so that data are comparable and can be used for risk management purposes.

9.4 Levels and patterns of contamination of food commodities

Data were submitted by Belgium, Canada, Japan, New Zealand, Poland and the USA and by the European Commission in a report containing data on Belgium, Denmark, Finland, France, Germany, Italy, The Netherlands, Norway, Sweden and the United Kingdom. In all countries in which a substantial number of samples had been analysed, the concentrations of PCDDs, PCDFs and coplanar PCBs in food were decreasing up to the late 1990s, but this decrease had slowed or was even partly reversed in some food categories in several countries owing to contamination of animal feed. For the present assessment of intake at the international level, only data collected after 1995 were considered.

As the Committee did not have access to the original analytical results, it was not possible to ascertain whether the results had been obtained by the lower- or upper-bound approach, and the concentrations used in the assessment were expressed as sums of congeners.

Insufficient individual data were available from most countries to allow construction of a full curve of the distribution of concentrations. Most data were submitted in an aggregated format. As recommended by a FAO/WHO workshop on assessing exposure to contaminants (WHO, 2000), aggregated data were weighted as a function of the number of initial samples and then used to obtain a weighted mean concentration of PCDDs, PCDFs and PCBs in six major food groups: meat and meat products, eggs, fish and fish products, milk and milk products, vegetables and vegetable products, and fats and oils. National data were aggregated by region or country (western Europe, Japan, New Zealand and North America) and are summarized in Table 45. Insufficient data were available for the rest of the world to permit a realistic estimate of the distribution of contaminants. The Committee recognized that there are significant differences within the food categories in Table 13 (p. 546), and that the data used in this analysis may not reflect the true mean for a food category. For example, the mean concentrations of PCDDs, PCDFs and coplanar PCBs and the rate of consumption vary considerably in different fish species, and it was not possible to determine if the mean represents the fish species most commonly consumed. However, the data received were not sufficient to allow an analysis that might account for such variation.

In a second step, a log-normal distribution of contaminants in foods was assumed, and a model of distribution was constructed from the weighted mean and a geometric standard deviation of 3 derived from the concentrations in six broad food groups. On the basis of these derived distributions, the percentiles of consumption were determined. The derived median values (50th percentiles) are presented in Table 13.

9.5 Estimated dietary intake

Because of the long half-lives of PCDDs, PCDFs and coplanar PCBs, their hazard to health can be estimated only after consideration of intake over a period of months. Short-term variations in PCDD, PCDF and coplanar PCB concentrations in foods have much less effect on overall intake than might be the case for other food contaminants.

The distribution of long-term mean intake in various populations was calculated by the following procedure:

  • The distributions of concentrations were constructed for various regions and food groups from the available data. The distributions were assumed to be log-normal.
  • Data on food consumption from the GEMS/Food regional diets and national surveys were used to estimate mean consumption of six major food groups in each diet. A log-normal distribution was constructed from these data with a geometric standard deviation of 1.3 extrapolated from the results of the food consumption survey in the Netherlands in order to account for inter-individual variation in consumption. The average contributions of the six basic food groups to total food consumption were derived for each diet.
  • The dietary intake of a particular population was assessed by combining the concentrations in food and food consumption distributions for that population with a Monte Carlo approach. In each Monte Carlo trial, the dietary intake was estimated by multiplying random values for food consumption and concentrations in various food groups. The concentrations were weighted according to the contribution of the food group to total food consumption. The estimates of intake were combined to form a distribution of long-term mean dietary intake for each population studied. The distributions are characterized by median and 90th percentile intake. Calculations were performed for the sum of the toxic equivalents of PCDDs and PCDFs and for the sum of coplanar PCBs separately, because the data on occurrence of PCBs were obtained independently.

The simulated intakes of PCDDs, PCDFs and coplanar PCBs in the GEMS/Food regional diets are presented in Table 45. These intakes are, however, likely to be overestimates, as the data on concentrations were derived from surveys (without random sampling) and from the GEMS/Food regional diets, which are based on data on food supply (apparent consumption), which are known to overestimate food consumption by at least 15%.

Table 45. Median and 90th percentile of estimated long-term intakes of toxic equivalents (pg/kg of body weight per month, assuming 60 kg of body weight) based on the GEMS/Foods regional diets

Source of data on concentrationsa

Source of data on food consumption

Intake of PCDDs/ PCDFs

Intake of coplanar PCBs

Median

90th percentile

Median

90th percentile

Western Europe

Europe

54

130

57

150

North America

Europe

68

160

14

35

New Zealand

Europe

18

36

10

22

Japan

Far East

7

15

7

19

a

For North America, the data on concentrations in vegetables in western Europe were used; for New Zealand, the data on concentrations in eggs in Japan were used.

More reliable estimates of intake (Table 46) were obtained by using national food consumption data rather than data on the food supply (apparent consumption) from the GEMS/Food regional diets. The simulated intakes presented in Table 46 are not strictly national estimates and are somewhat higher than the national estimates submitted by the European Commission.

Table 46. Median and 90th percentile of estimated long-term intakes of toxic equivalents (pg/kg of body weight per month, assuming 60 kg of body weight) based on national food consumption data

Source of data on concentrationsa

Source of data on food consumption

Intake of PCDDs/PCDFs

Intake of coplanar PCBs

Median

90th percentile

Median

90th percentile

North America

USA

42

100

9

25

Western Europe

France

40

94

47

130

Western Europe

Netherlands

33

81

30

82

Western Europe

United Kingdom

39

91

41

110

a For North America, the data on concentrations in vegetables in western Europe were used.

The calculated contributions of various food categories to the intake of PCDDs, PCDFs and coplanar PCBs showed that the largest fraction (> 70%) is from food of animal origin in both the GEMS/Food regional and national diets.

Information was lacking on both the quality of data and geographical represen-tativeness for some regions. More data are required on the occurrence of coplanar compounds in food products, particularly from geographical regions other than Europe, so that more representative estimates of intake can be made for all regions.

Breastfed infants have higher intakes of these compounds than bottle-fed infants or adults on a body-weight basis, although for only a small portion of their lives. Breast milk has beneficial effects, despite the contaminants present. WHO has therefore repeatedly evaluated the health significance of contamination of breast milk with coplanar compounds. WHO recommends and supports breastfeeding but has concluded that continued and enhanced efforts should be directed towards identifying and controlling environmental sources of these substances.

10. EVALUATION

In view of the long half-lives of PCDDs, PCDFs and coplanar PCBs, the Committee concluded that it would not be appropriate to establish an acute reference dose for these compounds.

The Committee concluded that a tolerable intake could be established for TCDD on the basis of the assumption that there is a threshold for all effects, including cancer. Carcinogenicity due to TCDD was not linked to mutagenicity or DNA binding, and it occurred at higher body burdens in animals than other toxic effects. The Committee concluded that the establishment of a tolerable intake based on effects other than cancer would also address any carcinogenic risk.

The studies listed in Table 44 were those considered by the Committee in choosing the lowest LOELs and NOELs for assessment of tolerable intake. The lowest LOEL and NOEL were provided by the studies of Faqi et al. (1998) and Ohsako et al. (2001), respectively. With the toxicokinetic conversions described in Table 43, these two studies indicate maternal body-burden LOELs and NOELs for effects on male rat offspring of 25 ng/kg bw and 13 ng/kg bw, respectively.

Background body burdens in laboratory animals

In the studies used to estimate body burden on the basis of the distribution of TCDD after multiple dosing, radiolabelled material was used. Therefore, the known background concentrations of TCDD and other PCDDs and PCDFs in the tissues of laboratory rodents resulting from traces of these compounds in rat feed were ignored. The Committee identified two studies that could be used to predict the body burdens of rats resulting from the presence of coplanar compounds in laboratory feed. These studies were mutually consistent and predicted that ‘unexposed’ laboratory rats had toxic equivalent body burdens of 3–12 ng/kg bw, depending on age. Thus, the maternal body burdens of TCDD seen in studies with radiolabelled material should be adjusted upwards by a minimum of 3 ng/kg bw to account for the background concentrations of unlabelled PCDDs and PCDFs. The maternal toxic equivalent body burden may still be underestimated, as 3 ng/kg bw was the minimum in the two studies, and in one of the studies coplanar PCBs were not included.

Addition of 3 ng/kg bw to the body burdens calculated from the linear model and the data in Table 43 resulted in estimated total toxic equivalent body burdens of 16 ng/kg bw for the NOEL and 28 ng/kg bw for the LOEL. These body burdens correspond to EHMIs of 240 and 420 pg/kg bw, respectively. Fitting the data in Table 43 into the power equation gave EHMIs of 330 pg/kg and 630 pg/kg, respectively.

Identification of safety factors

The safety factors considered in establishing acceptable levels of intake on the basis of the results of studies in laboratory animals usually include the following: a factor to convert a LOEL to a NOEL (if needed); a factor to extrapolate from animals to humans; and factors to account for inter-individual variations in susceptibility. Typically, factors of 10 have been used for extrapolation between species and to account for human variation in susceptibility, and a factor of 3–10 for extrapolating from a LOEL to a NOEL.

As a NOEL was identified for effects in the male offspring of rats, no factor for conversion from a NOEL to a LOEL was needed for the EHMI derived from the study described above (Faqi et al., 1998).

As concluded by the WHO consultation (van Leeuwen & Younes, 2000), use of body burdens to scale doses from studies in laboratory animals to equivalent human doses removes the need for safety factors to account for differences in toxicokinetics between animals and humans.

To account for interindividual differences in toxicokinetics among humans, a safety factor should be applied. The Committee noted that limited data were available on the toxicokinetics of TCDD in humans and considered that the default factor of 3.2 was appropriate.

The Committee observed that humans may be less sensitive than rats to some effects, but the conclusion is less certain for others, and it cannot be excluded that the most sensitive humans might be as sensitive to the adverse effects of TCDD as rats were in the pivotal studies. Therefore, the Committee concluded that no safety factor in either direction need be applied for differences in toxicodynamics among humans.

Use of a LOEL instead of a NOEL indicates the need for an additional safety factor. As the LOEL for the sensitive end-point was considered to be close to a NOEL and represented marginal effects, the Committee applied a factor of 3 to account for use of a LOEL instead of a NOEL. This resulted in an overall safety factor of 9.6 (3 Χ 3.2).

The Committee concluded that a total safety factor of 3.2 should be applied to the EHMI associated with the NOEL, and a total safety factor of 9.6 should be applied to the EHMI associated with the LOEL.

Tolerable intake

As stated in the discussion of toxicokinetics, the long half-lives of PCDDs, PCDFs and coplanar PCBs mean that each daily ingestion has a small or even a negligible effect on overall intake. In order to assess long- or short-term risks to health due to these substances, total or average intake should be assessed over months, and tolerable intake should be assessed over a period of at least 1 month. To encourage this view, the Committee decided to express the tolerable intake as a monthly value in the form of a provisional tolerable monthly intake (PTMI)2.

As shown in Table 47, use of the linear model to extrapolate the maternal body burden at the NOEL, obtained with a single dose, to that expected at multiple doses gives a EHMI of 237 pg/kg bw, which would be expected to result in a body burden that is lower than that which had effects in animals. The PTMI derived by application of the safety factor of 3.2 to this EHMI is 74 pg/kg bw.

Table 47. Summary of four calculations of PTMI

 

Linear model

Power model

NOEL

LOEL

NOEL

LOEL

Administered dose (ng/kg bw)

12.5a

 

12.5a

 

Maternal body burden (ng/kg bw)

7.6

25b

7.6

25b

Equivalent maternal body burden with long-term dosing (ng/kg bw)

13c

25c

19d

39d

Body burden from feed (ng/kg bw)

3

3

3

3

Total body burden (ng/kg bw)

16e

28e

22e

42e

EHMI (pg/kg bw per month)

237

423

330

630

Safety factor

3.2

9.6

3.2

9.6

PTMI (pg/kg bw per month)

74

44

103

66

a

Bolus dose (NOEL)

b

Target maternal body burden from repeated dosing (LOEL)

c

Assuming a linear relationship between fetal and maternal body burden (based on data in Table 43).

d

Assumes a non-linear relationship between fetal and maternal body burden (based on data in Table 43).

e

Assuming, for humans, a 7.6 year half-life and 50% uptake from food (see Eq. 1 on p. 615).

Similarly, as shown in Table 47, the PTMI derived by application of the safety factor of 9.6 to the EHMI derived from the study that provided the LOEL is 44 pg/kg bw. As also shown in Table 47, use of the power model to extrapolate the maternal body burden with single doses to multiple doses would result in PTMIs of 103 pg/kg bw for the NOEL and 66 pg/kg bw for the LOEL. The range of PTMIs derived from the two studies, with either the linear or the power model to extrapolate the maternal body burden with single to multiple doses, is 40–100 pg/kg bw per month. The Committee chose the mid-point of this range, 70 pg/kg bw per month, as the PTMI.

Furthermore, in accordance with the conclusions of the WHO consultation, the Committee concluded that this tolerable intake should be applied to intake of PCDDs, PCDFs and coplanar PCBs expressed as TEFs.

Comparison of PTMI with estimated intake from food

In the GEMS/Food regional diets, the range of estimated intake of toxic equivalents of PCDDs and PCDFs is 7–68 pg/kg bw per month at the median and 15–160 pg/kg bw per month at the 90th percentile of mean lifetime exposure, and those for coplanar PCBs are 7–57 pg/kg bw per month at the median and 19–150 pg/kg bw per month at the 90th percentile of consumption. The intakes estimated from national food consumption data were lower: 33–42 pg/kg bw per month at the median and 81–100 pg/kg bw per month at the 90th percentile for PCDDs and PCDFs, and 9–47 pg/kg bw per month at the median and 25–130 pg/kg bw per month at the 90th percentile for coplanar PCBs. Estimates could not be made for the sum of PCDDs, PCDFs and coplanar PCBs, because data on concentrations were submitted separately by countries.

The median and 90th percentile of the derived distribution of intakes were considered to describe long-term intake. A Monte Carlo calculation was used to predict these intakes for coplanar PCBs on the basis of two sets of distribution curves generated from information on mean concentrations in six major food groups and corresponding data on mean food consumption from several sources, by applying geometric standard deviations of 3 and 1.3 to the respective means. The geometric standard deviation for the food consumption curves accounted for long-term consumption patterns. As the mean intakes of the whole population tend not to change with the duration of a survey, use of mean consumer intakes to generate the curves for major food groups, rather than individual commodities, approximates the mean intakes of the whole population, as nearly all respondents were consumers.

Uncertainties

Several sources of uncertainty were identified in the data used to assess intake, which suggest that they are likely to be overestimates at both the median and the 90th percentile levels of consumption. Despite the uncertainties, the results suggest that a considerable fraction of the population will have long-term mean intake above the PTMI.

Furthermore, despite the large amount of information on toxicity, substantial uncertainties remain which should be considered in applying the risk assessment and in interpreting the estimates of intake of PCDDs, PCDFs and coplanar PCBs. The Committee used the overall data to identify a level of intake of coplanar compounds in food that represents no appreciable risk to humans. The safety assessment includes adjustment for a number of uncertainties, including estimates of TEFs within orders of magnitude in order to relate the potency of 28 relatively poorly studied compounds to that of one well-studied compound, TCDD. Moreover, the relative proportion of TCDD and the other 28 compounds varies; TCDD typically constitutes a small percentage of the total toxic equivalents in foods.

The PTMI is not a limit of toxicity and does not represent a boundary between safe intake and intake associated with a significant increase in body burden or risk. Long-term intakes slightly above the PTMI would not necessarily result in adverse health effects but would erode the safety factor built into the calculations of the PTMI. It is not possible, given current knowledge, to define the magnitude and duration of excess intake that would be associated with adverse health effects.

Effect of maximum limits on intake, risk and food availability

The concentrations of PCDDs, PCDFs and coplanar PCBs vary among foods. In establishing regulatory limits, the possible undesired consequences of their enforcement should be taken into account, such as reductions in the food supply. The Committee explored the theoretical effects of various maximum regulatory limits on compliance and on long-term average reduction of intake. On the basis of this analysis, the Committee concluded that, in order to achieve, for example, a 20% reduction in intake of coplanar compounds from food, the intake of a wide range of foods would have to be reduced by a similar percentage. This relationship exists because these contaminants are present at relatively high levels in major food types. Furthermore, in view of the half-lives of these compounds in humans, setting regulatory limits on the basis of the PTMI would have no discernible effect on body burdens for several years.

In contrast, long-term reductions could be gained by identifying and eliminating the routes by which these compounds pass from the environment into food supplies. The Committee was informed that studies of environmental concentrations over time in several countries suggest that measures to control emissions to the environment generally have had a substantial impact on both the amounts of PCDDs and PCDFs present in the environment and the body burdens of the general public.

10. REFERENCES

Abbott, B.D., Schmid, J.E. & Pitt, J.A. (1999) Adverse reproductive outcomes in the transgenic Ah receptor-deficient mouse. Toxicol. Appl. Pharmacol., 155, 62–70.

Abel, J., Li, W., Döhr, O., Vogel, C. & Donat, S. (1996) Dose–response relationship of cytochrome P4501b1 mRNA induction by 2,3,7,8-tetrachlorodibenzo-p-dioxin in livers of C57Bl/6J and DBA/2J mice. Arch. Toxicol., 70, 510–513.

Abernethy, D.J. & Boreiko, C.J. (1987) Promotion of C3H/T101/2 morphological transformation by polychlorinated dibenzo-p-dioxin isomers. Carcinogenesis, 8, 1485–1490.

Abernethy, D.J., Greenlee, W.F., Huband, J.C. & Boreiko, C.J. (1985) 2,3,7,8-Tetrachlorodi-benzo-p-dioxin (TCDD) promotes the transformation of C3H/10T1/2 cells. Carcinogenesis, 6, 651–653.

Abraham, K., Krowke, R. & Neubert, D. (1988) Pharmacokinetics and biological activity of 2,3,7,8-tetrachlorodibenzo-p-dioxin. 1. Dose-dependent tissue distribution and induction of hepatic ethoxyresorufin-O-deethylase in rats following a single injection. Arch. Toxicol., 62, 359–368.

Abraham, K., Wiesmüller, T., Brunner, H., Krowke, R., Hagenmaier, H. & Neubert, D. (1989) Absorption and tissue distribution of various polychlorinated dibenzo-p-dioxins and dibenzofurans (PCDDs and PCDFs) in the rat. Arch. Toxicol., 63, 193–202.

AEA Technology (1999) Compilation of EU Dioxin Exposure and Health Data, Report to the European Commission (DG Environment)

Agence française de Sécurité Sanitaire des Aliments (2001) [Individual and national survey of food consumption.], INCA 1999 Available at: <http://www.afssa.fr/dossiers/index.asp?id_dossier=2842> (in French).

Agency for Toxic Substances and Disease Registry (1998) Toxicological profile for chlorinated dibenzo-p-dioxins (update) (NTIS # PB/99/121998), Washington DC.

Ahlborg, U.G., Becking, G.C., Birnbaum, L.S., Brouwer, A., Derks, H.J.G.M., Feeley, M., Golor, G., Hanberg, A. Larsen, J.C. & Liem, A.K.D. (1994) Toxic equivalency factors for dioxin-like PCBs: Report on a WHO-ECEH and IPCS consultation. Chemosphere, 28, 1049–1067.

Alaluusua, S., Lukinmaa, P.-L., Vartiainen, T., Partanen, M., Torppa, J. & Tuomisto, J. (1996) Polychlorinated dibenzo-p-dioxins and dibenzofurans via mother's milk may cause developmental defects in the child's teeth. Environ. Toxicol. Pharmacol., 1, 193–197.

Alaluusua, S., Lukinmaa, P.L., Torppa, J., Tuomisto, J. & Vartianen, T. (1999) Developing teeth as biomarker of dioxin exposure. Lancet, 353, 206.

Alawi, M.A., Wichmann, H., Lorenz, W. & Bahadir, M. (1996) Dioxins and furans in the Jordanian environment. Part 2: Levels of PCDD and PCDF in human milk samples from Jordan. Chemosphere, 33, 2469–2474.

Alderfer, R., Sweeney, M., Fingerhut, M., Hornung, R., Wille, K. & Fidler, A. (1992) Measures of depressed mood in workers exposed to 2,3,7,8-tetrachlorodibenzo-p-dioxin. Chemosphere, 25, 247–250.

Allen, J.R., Van Miller, J.P. & Norback, D.H. (1975) Tissue distribution, excretion and biological effects of [14C]tetrachlordibenzo-p-dioxin in rats. Food Cosmet. Toxicol.,13, 501–505.

Allen, J.R., Barsotti, D.A., Van Miller, J.P., Abrahamson, L.J. & Lalich, J.J. (1977) Morphological changes in monkeys consuming a diet containing low levels of 2,3,7,8-tetrachlorodibenzo-p-dioxin. Food Cosmet. Toxicol., 15, 401–410.

Allen, J.R., Barsotti, D.A., Lambrecht, L.K. & Van Miller, J.P. (1979) Reproductive effects of halogenated aromatic hydrocarbons on nonhuman primates. Ann. N.Y. Acad. Sci., 320, 419–425.

Alsharif, N.Z., Schlueter, W.J. & Stohs, S.J. (1994) Stimulation of NADPH-dependent reactive oxygen species formation and DNA damage by 2,3,7,8-tetrachlorodibenzo-p-dioxin in rat peritoneal lavage cells. Arch. Environ. Contam. Toxicol., 26, 392–397.

American Public Health Association (1998) P450 reporter gene response to dioxin-like organic compounds. Method 8070. In: Standard Methods for the Examination of Water and Wastewater, 20th Ed., Washington DC, pp. 8–37.

American Society for Testing and Materials (1999) Standard guide E 1853M-98 for measuring the presence of planar organic compounds which induce CYP1A, reporter gene test systems. In: Annual Book of ASTM Standards, Biological Effects; Environmental Fate; Biotechnology; Pesticides, Vol. 11.05, Section 11, West Conshohocken, PA.

Andersen, M.E., Mills, J.J., Gargas, M.L., Kedderis, L., Birnbaum, L.S., Neubert, D. & Greenlee, W.L. (1993) Modelling receptor-mediated processes with dioxin: Implications for pharmacokinetics and risk assessment. Risk Anal., 13, 25–36.

Andersen, M.E., Eklund, C.R., Mills, J.J., Barton, H.A. & Birnbaum, L.S. (1997a) A multi-compartment geometric model of the liver in relation to regional induction of cytochrome P450s. Toxicol. Appl. Pharmacol., 144, 135–144

Andersen, M.E., Birnbaum, L.W.S., Barton, H.A. & Eklund, C.R. (1997b) Regional hepatic CYP1A1 and CYP1A2 induction with 2,3,7,8-tetrachlorodibenzo-p-dioxin evaluated with a multicompartmental geometric model of hepatic zonation. Toxicol. Appl. Pharmacol., 144, 145–155.

Andersen, M.E., Gargas, M.L., Hays, S.M., Clewell, H.J. & Paustenbach, D.J. (1997) Estimating minimally effective inducing doses (ED01s) of TCDD in livers of rats and humans using physiologically based pharmacokinetic modeling. Organohalogen Compd, 34, 311–316.

Andrae, U. & Greim, H. (1992) Initiation and promotion in thyroid carcinogenesis. In: Tissue Specific Toxicity: Biochemical Mechanisms, New York: Academic Press, pp. 71–93.

Arnold, D.L., Mes, J., Bryce, F., Karpinski, K., Bickis, M.G., Zawidzka, Z.Z. & Stapley, R. (1990) A pilot study on the effects of Aroclor 1254 ingestion by rhesus and cynomolgus monkeys as a model for human ingestion of PCBs. Food Chem. Toxicol., 28, 847–857.

Arnold, D.L., Nera, E.A., Stapley, R., Tolnai, G., Claman, P., Hayward, S., Tryphonas, H. & Bryce, F. (1996) Prevalence of endometriosis in rhesus (Macaca mulatta) monkeys ingesting PCB (Aroclor 1254): Review and evaluation. Fundam. Appl. Toxicol., 31, 42–55.

Assennato, G., Cervino, D., Emmet, E., Longo, G. & Merlo, F. (1989) Follow-up on subjects who developed chloracne following TCDD exposure of Seveso. Am. J. Ind. Med., 16, 119–125.

Axelson, O., Sundell, L., Andersson, K., Edling, C., Hogstedt, C. & Kling, H. (1980) Herbicide exposure and tumor mortality. An updated epidemiologic investigation on Swedish railroad workers. Scand. J. Work Environ. Health, 6, 73–79.

Aylward, L.L., Hays, S., Karch, N.J. & Paustenbach, D.J. (1996) Relative susceptibility of animals and humans to cancer hazard posed by 2,3,7,8-tetrachlorodibenzo-p-dioxin using internal measures of dose. Environ. Sci. Technol., 30, 3534–3543.

Badesha, J.S., Maliji, G. & Flaks, B. (1995) Immunotoxic effects of prolonged dietary exposure of male rats to 2,3,7,8-tetrachlorodibenzo-p-dioxin. Eur. J. Pharmacol., 293, 429–437.

Ballschmiter, K. & Zell, M. (1980) Analysis of polychlorinated biphenyls (PCB) by glass capillary gas chromatography. Fresenius Z. Anal. Chem., 302, 20–31.

Bandiera, S., Farrell, K., Mason, G.M., Bannister, R. & Safe, S. (1984) Comparative toxicities of the polychlorinated dibenzofuran (PCDF) and biphenyl (PCB) mixtures which persist in Yusho victims. Chemosphere, 13, 507–512.

Banks, Y.B. & Birnbaum, L.S. (1991) Absorption of 2,3,7,8-tetrachloro-p-dioxin (TCDD) after low dose dermal exposure. Toxicol. Appl. Pharmacol., 107, 302–310.

Barter, R.A. & Klaassen, C.D. (1992) UDP-glucuronosyltransferase induces reduced thyroid hormone levels in rats by an extrathyroidal mechanism. Toxicol. Appl. Pharmacol., 113, 36–42.

Bastomsky, C.H. (1974) Effects of a polychlorinated biphenyl mixture (Aroclor 1254) and DDT on biliary thyroxin excretion in rats. Endocrinology, 95, 1150.

Bastomsky, C.H. (1977) Enhanced thyroxin metabolism and high uptake goiters in rats after a single dose of 2,3,7,8-tetrachlorodibenzo-p-dioxin. Endocrinology, 101, 292–296.

Becher, H., Flesch-Janys, D., Kauppinen, T., Kogevinas, M., Steindorf, K. & Wahrendorf, J. (1996) Excess cancer mortality in German male workers exposed to phenoxy herbicides and dioxins. Cancer Causes Control, 7, 312–321.

Beck, H., Eckart, K., Mathar, W. & Wittkowski, R. (1989) PCDD and PCDF body burden from food intake in the FRG. Chemosphere, 18, 417–424.

Beetstra, J.B., van Engelen, J.G.M., Karel, P., van der Hoek, M., de Jong, M., Docter, R., Krenning, E.P., Hennemann, G., Brouwer, A. & Visser, T.J. (1991) Thyroxine and 3,3΄,5-triiodothyronine are glucuronidated in rat liver by different uridine diphosphate-glucuronyltransferases. Endocrinology, 128, 741.

Behnisch, P.A., Hosoe K. & Sakai S. (2001a) Bioanalytical screening methods for dioxins and dioxin-like compounds—A review of bioassay/biomarker technology. Environ. Int. (in press).

Behnisch, P.A., Hosoe, K. & Sakai, S. (2001b) Combinatorial bio/chemical analysis of dioxin and dioxin-like compounds in waste recycling, food, humans/wildlife and the environment. Environ. Int. (in press).

Behnisch, P.A., Allen, R., Anderson, J., Brouwer, A., Brown, D.J., Campbell, T.C., Guoyens, L., Harrison, R.O., Hoogenboom, R., van Overmeire, I. & Malisch, R. (2001c) Harmonized quality criteria for chemical and bioassay analyses of PCDDs/PCDFs in feed and food. Part 2: General considerations, bioassay methods submitted for Dioxin 2001. Organohalogen Compd (in press).

Belgian Federal Government. on line: http://www.dioxin.fgov.be

Belgian Federal Government (1999) Koninklijk besluit tot vaststelling van maximale gehaltes aan dioxines en polygechloreerde bifenylen in sommige voedingsmiddelen. Belg. Staatsblad, 3, 50495–50496 (in Dutch and French).

van den Berg, M., de Jongh, J. Poiger, H. & Olson, J.R. (1994) The toxicokinetics and metabolism of polychlorinated dibenzo-p-dioxins (PCDDs) and dibenzo-p-furans (PCDFs) and their relevance for toxicity. Crit. Rev. Toxicol., 24, 1–74.

van den Berg, M., Birnbaum, L., Bosveld, A.T.C., Brunström, B., Cook, P., Feeley, M., Giesy, J.P., Hanberg, A., Haegawa, R., Kennedy, S.W., Kubiak, T., Larsen, J.C., van Leeuwen, F.X.R., Liem, A.K.D., Nolt, C., Peterson, R.E., Poellinger, L., Safe, S., Schrenk, D., Tillitt, D., Tyslind, M., Younes, M., Waern, F. & Zacharewski, T. (1998) Toxic equivalency factors (TEFs) for PCBs, PCDDs, PCDFs for humans and wildlife. Environ. Health Perspectives, 106, 775–792.

van den Berg, M., Peterson, R.E. & Schrenk, D. (2000) Human risk assessment and TEFs. Food Addit. Contam., 17, 347–358.

Bertazzi, P.A. & di Domenico, A. (1994) Chemical, environmental and health aspects of the Seveso, Italy, accident. In: Schecter, A., ed., Dioxins and Health, New York, Plenum Press, pp. 587–632.

Bertazzi, P.A., Zocchetti, C., Pesatori, A.C., Guercilena, S. & Sanarico, M. (1989) Ten-year mortality study of the population involved in the Seveso incident in 1976. Am. J. Epidemiol., 129, 1187–200.

Bertazzi, P.A., Pesatori, A.C., Consonni, D., Tironi, A., Landi, M.T. & Zocchetti, C. (1993) Cancer incidence in a population accidentally exposed to 2,3,7,8-tetrachlorodibenzo-para-dioxin. Epidemiology, 4, 398–406.

Bertazzi, P.A., Zocchetti, C., Guercilena, S., Consonni, D., Tironi, A., Landi, M.T. & Pesatori, A.C. (1997) Dioxin exposure and cancer risk: A 15-year mortality study after the 'Seveso accident'. Epidemiology, 8, 646–652.

Bertazzi, P.A., Bernucci,I., Brambilla, G., Consonni, D. & Pesatori, A.C. (1998a) The Seveso studies on early and long-term effects of dioxin exposure: A review. Environ. Health Perspectives, 106, 625–633.

Bertazzi, P.A., Pesatori, A.C & Zocchetti, C. (1998b) Seveso. Dioxin: An example of environmental medicine. Epidemiologic data as guidelines for health programming. G. Ital. Med. Lav. Ergon., 20, 194–196.

Bertazzi, P.A., Consonni, D., Bachetti, S., Rubagotti, M., Baccarelli, A., Zocchetti, C. & Pesatori, A.C. (2001) Health effects of dioxin exposure: A 20-year mortality study. Am. J. Epidemiol. (in press).

van Birgelen, A.P.J.M., van der Kolk, J., Fase, K.M., Bol, I., Poiger, H., van den Berg, M. & Brouwer, A. (1994a) Toxic potency of 2,3,3΄,4,4΄,5΄-hexachlorobiphenyl relative to and in combination with 2,3,7,8-tetrachlorodibenzo-p-dioxin in a subchronic feeding study in the rat. Toxicol. Appl. Pharmacol., 126, 202–213.

van Birgelen, A.P.J.M., van der Kolk, J., Fase, K.M., Bol, I., Poiger, H., van den Berg, M. & Brouwer, A. (1994b) Toxic potency of 3,3΄,4,4΄,5΄-pentachlorobiphenyl relative to and in combination with 2,3,7,8-tetrachlorodibenzo-p-dioxin in a subchronic feeding study in the rat. Toxicol. Appl. Pharmacol., 127, 209–221.

van Birgelen, A.P.J.M., van der Kolk, J., Fase, K.M., Bol, I., Poiger, H., Brouwer, A. & van den Berg, M. (1995a) Subchronic dose–response study of 2,3,7,8-tetrachlorodibenzo-p-dioxin in female Sprague-Dawley rats. Toxicol. Appl. Pharmacol., 132, 1–13.

van Birgelen, A.P.J.M., Smit, E.A., Kampen, I.M., Groeneveld, C.N., Fase, K.M., van der Kolk, J., Poiger, H., van den Berg, M., Koeman, J.H. & Brouwer, A. (1995b) Subchronic effects of 2,3,7,8-TCDD or PCBs on thyroid hormone metabolism: Use in risk assessment. Eur. J. Pharmacol., 293, 77–85.

van Birgelen, A.P.J.M., Fase, K.M., van der Kolk, J., Poiger, Brouwer, A., Seinen, W. & van den Berg, M. (1996a) Synergistic effect of 2,2΄,4,4΄,5,5΄-hexachlorobiphenyl and 2,3,7,8-tetrachlorodibenzo-p-dioxin on hepatic porphyrin levels in the rat. Environ. Health Perspectives, 104, 550–557.

van Birgelen, A.P.J.M., DeVito, M.J., Akins, J.M., Ross, D.G., Diliberto, J.J. & Birnbaum, L.S. (1996b) Relative potencies of polychlorinated dibenzo-p-dioxins, dibenzofurans, and biphenyls derived from hepatic porphyrin accumulation in mice. Toxicol. Appl. Pharmacol., 138, 98–109.

Birnbaum, L.S. (1985) The role of structure in the disposition of halogenated aromatic xenobiotics. Environ. Health Perspectives, 61, 11–20.

Birnbaum, L.S. (1986) Distribution and excretion of 2,3,7,8-tetrachlorodibenzo-p-dioxin in congenic strains of mice which differ at the Ah-locus. Drug Metab. Disposition, 14, 34–40.

Birnbaum, L. (1998) Developmental effects of dioxins. In: Korach, K.S., ed., Reproductive and Developmental Toxicology, New York: Marcel Dekker, pp. 87–112.

Birnbaum, L.S. & Couture, L.A. (1988) Disposition of octachlorodibenzo-p-dioxin (OCDD) in male rats. Toxicol. Appl. Pharmacol., 93, 22–30.

Birnbaum, L.S. & Tuomisto, J. (2000) Non-carcinogenic effects of TCDD in animals. Food Addit. Contam., 17, 275–288.

Birnbaum, L.S., Decad, G.M. & Matthews, H.B. (1980) Disposition and excretion of 2,3,7,8-tetrachlorodibenzofuran in the rat. Toxicol. Appl. Pharmacol., 55, 342–352.

Birnbaum L., Harris, M.W., Miller, C.P., Pratt, R.M. & Lamb, J.C. (1986) Synergistic interaction of 2,3,7,8-tetrachlorodibenzo-p-dioxin and hydrocortisone in the induction of cleft palate in mice. Teratology, 33, 29–35.

Birnbaum, L.S., Harris, M.W., Stocking, L.M., Clark, A.M. & Morrisey, R.E. (1989) Retinoic acid and 2,3,7,8-tetrachlorodibenzo-p-dioxin selectively enhance teratogenesis in C57BL/6N mice. Toxicol. Appl. Pharmacol., 98, 487–500.

Birnbaum, L.S., McDonald, M.M., Blair, P.C., Clark, A.M. & Harris, M.W. (1990) Differential toxicity of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in C57BL/6J mice congenic at the Ah locus. Fundam. Appl. Toxicol., 15, 186–200.

Birnbaum, L.S., Ross, D.G. & DeVito, M.J. (1995) Dose–response relationships for EROD induction in liver, lung and skin for dioxin and dibenzofurans. Organohalogen Compd, 25, 237–240.

Blair, A., Grauman, D.J., Lubin, J.H. & Fraumeni, J.F., Jr (1983) Lung cancer and other causes of death among licensed pesticide applicators. J. Natl Cancer Inst., 71, 31–37.

Blaylock, B.L., Holladay, S.D., Comment, C.E., Heindel, J.J. & Luster, M.I. (1992) Exposure to tetrachlorodibenzo-p-dioxin (TCDD) alters fetal thymocyte maturation. Toxicol. Appl. Pharmacol., 112, 207–213.

Bleiberg, J., Wallen, M., Brodkin, R. & Applebaum, I.L. (1964) Industrially acquired porphyria. Arch. Dermatol., 89, 793–797.

Boersma, E.R., Lanting, C.I., Patandin, S., Weisglas-Kuperus, N., Touwen, B.C.L. & Sauer, P.J. (2000) Effects of perinatal exposure to background levels of PCBs and dioxins on the child's PCB body burden and on neurologic and cognitive development during the first 42 months of life. In: Risk Assessment in the Food Chain of Children (Nestlé Nutrition Workshop Series 44), pp. 47–60.

Bois, F.Y. & Eskenazi, B. (1994) Possible risk of endometriosis for Seveso, Italy, residents: An assessment of exposure to dioxin. Environ. Health Perspectives, 102, 476–477.

Bond, G.G., McLaren, E.A., Lipps, T.E. & Cook, R.R. (1989) Update of mortality among chemical workers with potential exposure to the higher chlorinated dioxins. J. Occup. Med., 31, 121–123.

Bovee, T.F.H., Hoogenboom, L.A.P., Hamers, A.R.M., Traag, W.A., Zuidema, T., Aarts, J.M.M.G., Brouwer, A. & Kuiper, H.A. (1998) Validation and use of the CALUX-bioassay for the determination of dioxins and PCBs in bovine milk. Food Addit. Contam., 15, 863–875.

Bowman, R.E., Schantz, S.L., Weerasinghe, N.C., Gross, M.L. & Barsotti, D.A. (1989) Chronic dietary intake of 2,3,7,8-tetrachlorodibenzo-p-dioxin at 5 or 25 parts per trillion in the monkey, TCDD kinetics and dose–effect estimate of reproductive toxicity. Chemosphere, 18, 243–252.

Brewster, D.W. & Birnbaum, L.S. (1987) Disposition and excretion of 2,3,4,7,8-pentachlorodi-benzofuran in the rat. Toxicol. Appl. Pharmacol., 90, 243–252.

Broeckaert, F. & Bernard, A. (2000) Polychlorinated Biphenyls (PCBs) and Dioxins (PCDD/Fs) in the Belgian Contaminated Food Chain: Sources, Profiles and Correlations. Final Report, Brussels: Federal Ministry of Agriculture.

Brouwer, A. & van den Berg, K.J. (1986) Binding of a metabolite of 3,4,3΄,4΄-tetrachlorobiphenyl to tranthyretin reduces serum vitamin A transport by inhibiting the formation of the protein complex carrying both retinol and thyroxin. Toxicol. Appl. Pharmacol., 85, 301.

Brouwer, A., Morse, D.C., Lans, M.C., Schuur, A.G., Murk, A.J., Klasson-Wehler E., Bergman, A. & Visser, T.J. (1998) Interactions of persistent environmental organohalogens with the thyroid hormone system, mechanisms and possible consequences for animal and human health. Toxicol. Ind. Health, 14, 59–84.

Brown, J.F., Lawton, R.W. & Ross, M.R. (1989) Persistence of PCB congeners in capacitor workers and Yusho patients. Chemosphere, 19, 829–834.

Brucker-Davis, F., (1998) Effects of environmental synthetic chemicals on thyroid function. Thyroid, 8, 827–856.

Bruner-Tran, K., Rier, S.E., Eisenberg, E. & Osteen, K.G. (1999) The potential role of environmental toxins in the pathophysiology of endometriosis. Gynecol. Obstet. Invest., 48 (Suppl. 1), 45–56.

Buckley-Kedderis, L.B., Andersen, M.E. & Birnbaum, L.S. (1993) Effect of dose, time, and pretreatment on the biliary excretion and distribution of 2,3,7,8-yetrachlorodibenzo-p-dioxin in the rat. Fundam. Appl. Toxicol., 121, 87–98.

Bueno de Mesquita, B.H., Doornbos, G., van der Deirdre, A.M., Kogevinas, M. & Winkelmann, R. (1993) Occupational exposure to phenoxy herbicides and chlorophenols and cancer mortality in The Netherlands. Am. J. Ind. Med., 23, 289–300.

Buhler, F., Schmid, P. & Schlatter, C.H. (1988) Kinetics of PCB elimination in man. Chemosphere, 17, 1717–1726.

Bund/Länder-Arbeitsgruppe Dioxine (1993) 2. Bericht 'Umweltpolitik', Bonn, Bundesministerium für Umwelt, Naturschutz und Reaktorsicherheit.

Bund/Länder-Arbeitsgruppe Dioxine (2001) 3. Bericht: Dioxin-Referenzmessprogramm. Bonn, Bundesministerium für Umwelt, Naturschutz und Reaktorsicherheit.

Burleson, G.R., Lebrec, H., Yang, Y.G., Ibanes, J.D., Pennington, K.N. & Birnbaum, L.S. (1996) Effect of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) on influenza virus host resistance in mice. Fundam. Appl. Toxicol., 29, 40–47.

Byrd, D.M. & Cothern, C.R. (2000) Introduction to Risk Analysis, Rockville MD: Government Institutes Publishers

Calvert, G.M., Sweeney, M.H., Fingerhut, M.A., Hornung, R.W. & Halperin, W.E. (1994) Evaluation of porphyria cutanea tarda in US workers exposed to 2,3,7,8-tetrachlorodibenzo-p-dioxin. Am. J. Ind. Med., 25, 559–571.

Calvert , G.M., Wille, K.K., Sweeney, M.H., Fingerhut, M.A. & Halperin, W.E. (1996) Evaluation of serum lipid concentrations among US workers exposed to 2,3,7,8-tetrachlorodibenzo-p-dioxin. Arch. Environ. Health, 51, 100–107.

Calvert, G.M., Sweeney, M.H., Deddens, J. & Wall, D.K. (1999) An evaluation of diabetes mellitus, serum glucose, and thyroid function among US workers exposed to 2,3,7,8-tetrachlorodibenzo-p-dioxin. Occup. Environ. Med., 56, 270–276.

Carrier, G., Brunet, R.C. & Brodeur, J. (1995a) Modeling of the toxicokinetics of polychlorinated dibenzo-p-dioxins and dibenzofurans in mammalians, including humans. I. Nonlinear distribution of PCDD/PCDF body burden between liver and adipose tissue. Toxicol. Appl. Pharmacol., 131, 253–266.

Carrier, G., Brunet, R.C. & Brodeur, J. (1995b) Modeling of the toxicokinetics of polychlorinated dibenzo-p-dioxins and dibenzofurans in mammalians, including humans. II. Kinetics of absorption and disposition of PCDDs/PCDFs. Toxicol. Appl. Pharmacol., 131, 267–276.

Centers for Disease Control (1988) Vietnam experience study, health status of Vietnam veterans, reproductive outcomes and child health. J. Am. Med. Assoc., 259, 2715–2719.

Chapman, D.E. & Schiller, C.M. (1985) Dose-related effects of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in C57BL/6J and DBA/2J mice. Toxicol. Appl. Pharmacol., 78, 147–157.

Chemisches Landes- und Staatliches Veterinäruntersuchungsamt Münster (1996) Final report, Laborvergleichsuntersuchung PCDD/F in Kuhmilch, Münster.

Chen, P.H., Luo, M.L. & Wong, C.K. (1982) Comparative rates of elimination of some individual polychlorinated biphenyls from the blood of PCB-poisoned patients in Taiwan. Food Chem. Toxicol., 20, 417–425.

Chen, P.H., Wong, C.-K., Rappe, C. & Nygren, M. (1985) Polychlorinated biphenyls, dibenzofurans and quaterphenyls in toxic rice-bran oil and in the blood and tissues of patients with PCB poisoning (Yu-cheng) in Taiwan. Environ. Health Perspectives, 59, 59–65.

Chen, L.-C., Berberian, I., Koch, B., Mercier, M., Azais-Braesco, V., Glauert, H.P., Chow, C.K. & Robertson, L.W., (1992a) Polychlorinated and polybrominated biphenyl congeners and retinoid levels in rat tissues: Structure–activity relationships. Toxicol. Appl. Pharmacol., 114, 47–55.

Chen, Y.-C.J., Guo, Y.-L., Hsu, C.-C. & Rogan, W.J. (1992b) Cognitive development of Yu-cheng ('oil disease') children prenatally exposed to heat-degraded PCBs. J. Am. Med. Assoc., 268, 3213–3218.

Clark, D.A., Gauldie, J., Szewczuk, M.R. & Sweeney, G. (1981) Enhanced suppressor cell activity as a mechanism of immunosuppression by 2,3,7,8-tetrachlorodibenzo-p-dioxin. Proc. Soc. Exp. Biol. Med., 168, 290–299.

Clark, D.A., Sweeney, G., Safe, S., Hancock, E., Kilburn, D.G. & Gauldie, J. (1983) Cellular and genetic basis for suppression of cytotoxic T cell generation by haloaromatic hydro-carbons. Immunopharmacology, 6, 143–153.

Clemons, J.H., Dixon, D.G. & Bois, N.C. (1995) Toxicant equivalent factors for selected PCBs measured by a fish, RTL-W1 and a rat, H411E, liver cell line. Organohalogen Compd, 26, 287–291.

Codex Alimentarius (2001) 33rd meeting of the Codex Committee on Food Additives and Contaminants (CCFAC) in March 2001 (http://www.codexalimentarius.net/Reports.htm).

Coenraads, P.J., Olie, K. & Tang, N.J. (1999) Blood lipid concentrations of dioxins and dibenzofurans causing chloracne. Br. J. Dermatol., 141, 694–697.

Coggon, D., Pannett, B., Winter, P.D., Acheson, E.D. & Bonsall, J. (1986) Mortality of workers exposed to 2 methyl-4 chlorophenoxyacetic acid. Scand. J. Work Environ. Health, 12, 448–454.

Coggon, D., Pannett, B. & Winter, P. (1991) Mortality and incidence of cancer at four factories making phenoxy herbicides. Br. J. Ind. Med., 48, 173–178.

Collins, W.T. & Capen, C.C. (1980) Ultrastructural and functional alterations of the rat thyroid gland produced by polychlorinated biphenyls compared with iodide excess and deficiency and thyrotropin and thyroxine administration. Virchows Arch. B Cell Pathol., 33, 213.

Collins, J.J., Strauss, M.E., Levinskas, G.J. & Conner, P.R. (1993) The mortality experience of workers exposed to 2,3,7,8-tetrachlorodibenzo-p-dioxin in a trichlorophenol process accident. Epidemiology, 4, 7–13.

Commission of the European Union (1996) Improvement of Knowledge of Food Consumption with a View to Protection of Public Health by Means of Exchanges and Collaboration between Database Managers (EUR 17528 EN), Brussels.

Commission of the European Union (1996) Certification of the Contents (Mass Fractions) of Five Polychlorinated-p-dioxines and Six Polychlorinated Furans in Milk Powder (EUR 16999 EN), Brussels: BCR Information, Reference Materials.

Commission of the European Union (1999) Compilation of EU dioxin exposure and health data. Available at: http://europa.eu.int/comm/environment/dioxin/index.htm.

Commission of the European Union (2000a) Reports on tasks for scientific cooperation, Task 3.2.5: Assessment of dietary intake of dioxins and related PCBs by the population of European Union, Member States, 7 June 2000. http://europa.eu.int/comm/dgs/health_consumer/library/pub/ pub08_en.pdf

Commission of the European Union (2000b) Opinion of the Scientific Committee on Food on the risk assessment of dioxins and dioxin-like PCBs in food, 22 November 2000. Brussels: Health and Consumer Protection Directorate-General, Scientific Co-operation and Networks. Available at: http://europa.eu.int/comm/food/fs/sc/scf/outcome_en.html.

Commission of the European Union (2000c) Opinion of the Scientific Committee on Animal Nutrition on the dioxin contamination of feedingstuffs and their contribution to the contamination of food of animal origin, 6 November 2000. Available at: http://europa.eu.int/comm/food/fs/sc/scan/ out55_en.pdf.

Cooke, M., Clark, G.C., Goeyens, L. & Baeyens, W. (2000) Environmental bioanalysis of dioxin. Today’s Chemist at Work, July, pp. 34–39.

Cordier, S., Le, T.B., Verger, P., Bard, D., Le, C.D., Larouze, B., Dazza, M.C., Hoang, T.Q. & Abenhaim, L. (1993) Viral infections and chemical exposures as risk factors for hepatocellular carcinoma in Vietnam. Int. J. Cancer, 55, 196–201.

Couture, L.A., Elwell, M.R. & Birnbaum, L.S. (1988) Dioxin-like effects observed in male rats following exposure to octachlorodibenzo-p-dioxin during a 13 week study. Toxicol. Appl. Pharmacol., 93, 31–46.

Cuijpers, C.E.J., Zeilmaker, G.W., van der Molen, W. & Slob, E.L. (1997) Developments in infant exposure to dioxins, furans and PCBs in breast milk and potential health consequences, in The Netherlands (RIVM report No. 529102007), Bilthoven: National Institute of Public Health and the Environment.

Cummings, A.M., Metcalf, J.L. & Birnbaum, L. (1996) Promotion of endometriosis by 2,3,7,8- tetrachlorodibenzo-p-dioxin in rats and mice: Time–dose dependence and species comparison. Toxicol. Appl. Pharmacol., 138, 131–139.

Dahl, P., Lindström, G., Wiberg, K. & Rappe, C. (1995) Absorption of polychlorinated biphenyls, dibenzo-p-dioxins and dibenzofurans by breast-fed infants. , 30, 2297–2306.

Davis, D.L., Gottlieb, M.B. & Stampnitzky, J.R., (1998) Reduced sex ratio of male to female births in several industrial countries. A sentinel health indicator? J. Am. Med. Assoc., 279,1018–1023.

Davis, D. & Safe, S. (1988) Immunosuppressive activities of polychlorinated dibenzofuran congeners: Quantitative structure–activity relationships and interactive effects. Toxicol. Appl. Pharmacol., 94, 141–149.

Decad, G.M., Birnbaum, L.S. & Matthews, H.B. (1981) Distribution and excretion of 2,3,7,8-tetrachlorodibenzofuran in C57BL/6J and DBA/2J mice. Toxicol. Appl. Pharmacol., 59, 564–573.

DeCaprio, A.P., McMartin, D.N., O’Keefe, P.W., Rej, R., Silkworth, J.B. & Kaminsky, L.S. (1989) Subchronic oral toxicity of 2,3,7,8-tetrachlorodibenzo-p-dioxin in the guinea pig: Comparisons with a PCB-containing transformer fluid pyrolysate. Fundam. Appl. Toxicol., 6, 454–463.

De Jongh, J., Wondergem, F., Seinen, W. & van den Berg, M. (1993a) Toxicokinetic interactions between chlorinated aromatic hydrocarbons in the liver of the C57BL/6J mouse: I. Polychlorinated biphenyls (PCBs). Arch. Toxicol., 67, 453–460.

De Jongh, J., Nieboer, R., Schroders, I, Seinen, W. & van den Berg, M. (1993b) Toxicokinetic mixture interactions between chlorinated aromatic hydrocarbons in the liver of the C57BL/6J mouse: 2. Polychlorinated dibenzo-p-dioxins (PCDDs), dibenzofurans (PCDFs) and biphenyls (PCBs). Arch. Toxicol., 67, 598–604.

DeJongh, J., DeVito, M., Nieboer, R., Birnbaum, L. & van de Berg, M. (1995) Induction of cytochrome P450 isoenzymes after toxicokinetic interactions between 2,3,7,8-tetrachlordibenzo-p-dioxin and 2,2΄,4,4΄,5,5΄-hexachlorobiphenyl in the liver of the mouse. Fundam. Appl. Toxicol., 25, 264–270.

Della Porta, G., Dragani, T.A. & Sozzi, G. (1987) Carcinogenic effects of infantile and long-term 2,3,7,8-tetrachlorodibenzo-p-dioxin treatment in the mouse. Tumori, 73, 99–107.

Department of Agriculture (1997) Data tables: Results from USDA’s 1996 continuing survey of food intakes by individuals and 1996 diet and health knowledge survey. Agricultural Research Service, ARS Food Surveys Research Group. Available at: <http://www.barc.usda.gov/bhnrc/foodsurvey/home.htm>.

Deutsche Forschungsgemeinschaft (1988) Polychlorierte Biphenyle, Bestandsaufnahme über Analytik, Vorkommen, Kinetik und Toxikologie, Mitteilung XIII der Senatskommission zur Prüfung von Rückständen in Lebensmitteln, Weinheim: VCH Verlagsgesellschaft.

DeVito, M.J. & Birnbaum, L.S. (1995) The importance of pharmacokinetics in determining the relative potency of 2,3,7,8-tetrachlorodibenzo-p-dioxin and 2,3,7,8-tetrachlorodibenzofuran. Fundam. Appl. Toxicol., 24, 145–148.

DeVito, M.J., Maier, W.E., Diliberto, J.J. & Birnbaum, L.S. (1992) Comparative ability of various PCBs, PCDFs and TCDD to induce cytochrome P450 1A1 and 1A2 activity following 4 weeks of treatment. Fundam. Appl. Toxicol., 20, 125–130.

DeVito, M.J., Maier, W.E., Diliberto, J.J. & Birnbaum, L.S. (1993) Comparative ability of various PCBs, PCDFs, and TCDD to induce cytochrome P450 1A1 and 1A2 activity following 4 weeks of treatment. Fundam. Appl. Toxicol., 20, 125–130.

DeVito, M.J., Ma, X., Babish, J.G., Menache, M. & Birnbaum, L.S. (1994) Dose–response relationships in mice following subchronic exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin: CYP1A1, CYP1A2, estrogen receptor, and protein tyrosine phosphorylation. Toxicol. Appl. Pharmacol., 124, 82–90.

DeVito, M.J., Birnbaum, L.S., Farland, W.H. & Gasiewicz, T.A. (1995) Comparisons of estimated human body burdens of dioxinlike chemicals and TCDD body burdens in experimentally exposed animals. Environ. Health Perspectives, 103, 820–831.

DeVito, M.J., Beebe, L.E., Menache, M. & Birnbaum, L.S. (1996) Relationship between CYP1A enzyme activities and protein levels in rats treated with 2,3,7,8-tetrachlorodibenzo-p-dioxin. J. Toxicol. Environ. Health, 47, 379–394.

DeVito, M.J., Ross, D.G., Dupuy, A.E., Ferrario, J., McDaniel, D. & Birnbaum, L.S. (1998) Dose–response relationships for disposition and hepatic sequestration of polyhalogenated dibenzo-p-dioxins, dibenzofurans and biphenyls following subchronic treatment in mice. Toxicol. Sci., 46, 223–234.

DeVito, M.J., Menache, M.G., Diliberto, J.J., Ross, D.G. & Birnbaum, L.S. (2000) Dose–response relationships for induction of CYP1A1 and CYP1A2 enzyme activity in liver, lung, and skin in female mice following subchronic exposure to polychlorinated biphenyls. Toxicol. Appl. Pharmacol., 167, 157–172.

Diliberto, J.J., Akubue, P.I., Luebke, R.W. & Birnbaum, L.S. (1995) Dose–response relationships of tissue distribution and induction of CYP1A1 and CYP1A2 enzymatic activities following acute exposure to 2,3,7,8-tetrachlorodibnzo-p-dioxin in mice. Toxicol. Appl. Pharmacol., 130, 197–208.

Diliberto, J.J., Jackson, J.A. & Birnbaum, L.S. (1996) Comparison of 2,3,7,8-tetrachlorodibnzo-p-dioxin (TCDD) Disposition following pulmonary, oral, dermal and parenteral exposures to rats. Toxicol. Appl. Pharmacol., 138, 158–168.

Diliberto, J.J., Burgin, D. & Birnbaum, L.S. (1997) Role of CYP1A2 in hepatic sequestration of dioxin: Studies using CYP1A2 knock-out mice. Biochem. Biophys. Res. Commun., 236, 431–433.

Diliberto,, J.J., DeVito, M., Ross, D.G. & Birnbaum, L.S. (1998) Time-course and dose–response relationships of sub-chronic dosing with [3H]-2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) on dosimetry and CYP1A1 and CYP1A2 activities in mice. Organohalogen Compd, 37, 381–384.

Diliberto, J.J., Burgin, D.E. & Birnbaum, L.S. (1999) Effects of CYP1A2 on disposition of 2,3,7,8-tetrachlorodibenzo-p-dioxin, 2,3,4,7,8-pentachlorodibenzofuran, and 2,2΄,4,4΄,5,5΄-hexaclorobiphenyl in CYP1A2 knockout and parental (C57BL/6N and 129/Sv) strains of mice. Toxicol. Appl. Pharmacol., 159, 52–64.

Dooley, R.K., Morris, D.L. & Holsapple, M.P. (1990) Elucidation of cellular targets responsible for tetrachlorodibenzo-p-dioxin (TCDD)-induced suppression of antibody response. 2. Role of the T lymphocyte. Immunopharmacology, 19, 47–58.

Dragnev, K.H., Beebe, L.E., Jones, C.R., Fox, S.D., Thomas, P.E., Nims, R.W. & Lubet, R.A. (1994) Subchonic dietary exposure to Aroclor 1254 in rats: Accumulation of PCBs in liver, blood, and adipose tissue and its relationship to induction of various hepatic drug-metabolising enzymes. Toxicol. Appl. Pharmacol., 125, 111–122.

Egeland, G.M., Sweeney, M.H., Fingerhut, M.A., Wille, K.K., Schnorr, T.M. & Halperin, W.E. (1994) Total serum testosterone and gonadotropins in workers exposed to dioxin. Am. J. Epidemiol., 139, 272–281.

Eldridge, S.R., Gould, M.N. & Butterworth, B.E. (1992) Genotoxicity of environmental agents in human mammary epithelial cellsAbbott, B.D., Schmid, J.E. & Pitt, J.A. (1999) Adverse reproductive outcomes in the transgenic Ah receptor-deficient mouse. Toxicol. Appl. Pharmacol., 155, 62–70.

Eljarrat, E., Caixach, J. & Rivera, J. (2000) Determination of PCDDs and PCDFs in different animal feed ingredients. Organohalogen Compd, 47, 330–333.

Environment Agency of Japan (1999) Report on Tolerable Daily Intake (TDI) of Dioxins and Related Compounds, Tokyo

Environmental Protection Agency (1992) Toxicity identification evaluations: Characterization of chronically toxic effluents, Phase 1 (EPA/600/6-91-005F), Duluth, MN: National Effluent Toxicity Assessment Center.

Environmental Protection Agency (2000a) Health assessment of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) and related compounds. Draft 9/18/2000. http://www.epa.gov/ncea/pdfs/dioxin/dioxreass.htm.

Environmental Protection Agency (2000b) Chlorinated dioxins and furans in the general US population: NHATS FY87 results (EPA-0560/5-91-003), Washington DC: Office of Toxic Substances.

Environmental Protection Agency (2000c) Method 4425: Screening extracts of environmental samples for planar organic compounds (PAHS, PCBS, PCDDS/PCDFS) by a reporter gene on a human cell line, Office of Solid Waste, SW 846 Methods, Update IVB, November, 2000.

Environmental Protection Agency (2000d) Workshop on the application of 2,3,7,8-TCDD toxicity equivalency factors to fish and wild-life (in review).

Fahrig, R. (1993) Genetic effects of dioxins on the spot test with mice. Environ. Health Perspectives, 101 (Suppl. 3), 257–261.

Faith, R.E. & Moore, J.A. (1977) Impairment of thymus-dependent immune function by exposure of the developing immune system to 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD). J. Toxicol. Environ. Health, 3, 451–464.

Faqi, A.S. & Chahoud, I. (1998) Antiestrogenic effects of low doses of 2,3,7,8-TCDD in offspring of female rats exposed throughout pregnancy and lactation. Bull. Environ. Contam. Toxicol., 61, 462–469.

Faqi, A.S., Dalsenter, P.R., Merker, H.J. & Chahoud, I. (1998) Reproductive toxicity and tissue concentrations of low doses of 2,3,7,8-tetrachlorodibenzo-p-dioxin in male offspring rats exposed throughout pregnancy and lactation. Toxicol. Appl. Pharmacol., 150, 383–392.

Fattore, E., Trossvik, C. & Häkansson, H. (2000) Relative potency values derived from hepatic vitamin A reduction in male and female Sprague-Dawley rats following subchronic dietary exposure to individual polychlorinated dibenzo-p-dioxin and dibenzofuran congeners and a mixture thereof. Toxicol. Appl. Pharmacol., 165, 184–194.

Feeley, M. & Brouwer, A. (2000) Health risks to infants from exposure to PCBs PCDDs and PCDFs. Food Addit. Contam., 17, 325–333.

Ferrario, J., McDaniel, D. & Byrne, C. (1999) The isomer distribution and congener profile of polychlorinated dibenzo-p-dioxins (PCDDs) in ball clay from the Mississippi Embayment (Sledge, Mississippi). Organohalogen Compd, 40, 95–99.

Filippini, G., Bordo, B., Crenna, P., Massetto, N., Musicco, M. & Boeri, R. (1981) Relationship between clinical and electrophysiological findings and indicators of heavy exposure to 2,3,7,8-tetrachlorodibenzo-dioxin. Scand. J. Work Environ. Health, 7, 257–262.

Fingerhut, M., Halperin, W., Marlow, D., et al. (1990) Mortality among US workers employed in the production of chemicals contaminated with 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) (NIOSH Final Report PB91125971), Springfield, VA: National Technical Information Service.

Fingerhut, M.A., Halperin, W.E., Marlow, D.A., Piacitelli, L.A., Honchar, P.A., Sweeney, M.H., Greife, A.L., Dill, P.A., Steenland, K. & Suruda, A.J. (1991) Cancer mortality in workers exposed to 2,3,7,8-tetrachlorodibenzo-p-dioxin. New Engl. J. Med., 324, 212–218.

Fitzgerald, E.F., Standfast, S.J., Youngblood, L.G., Melius, J.M. & Janerich, D.T. (1986) Assessing the health effects of potential exposure to PCBs, dioxins, and furans from electrical transformer fires: the Binghamton State Office Building medical surveillance program. Arch. Environ. Health, 41, 368–376.

Fitzgerald, E.F., Weinstein, A.L., Youngblood, L.G., Standfast, S.J. & Melius, J.M. (1989) Health effects three years after potential exposure to the toxic contaminants of an electrical transformer fire. Arch. Environ. Health, 44, 214–221.

Flaws, J.A., Sommer, R.J., Silbergeld, E.K., Peterson, R.E. & Hirshfield, A.N. (1997) In utero and lactational exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) induces genital dysmorphogenesis in the female rat. Toxicol. Appl. Pharmacol., 147, 351–362.

Flesch-Janys, D. (1998) Analyses of exposure to polychlorinated dibenzo-p-dioxins, furans and hexachlorocyclohexane and different health outcomes in a cohort of former herbicide-producing workers in Hamburg, Germany. Teratog. Carcinog. Mutag., 17, 257–264.

Flesch-Janys, D., Gurn, P., Jung, D., Konietzko, J. & Päpke, O. (1994) First results of an investigation of the elimination of polychlorinated dibenzo-p-dioxins and dibenzo-p-furans (PCDD/F) in occupationally exposed persons. Organohalogen Compd, 21, 93–99.

Flesch-Janys, D., Berger, J., Konietzko, J., Gurn, P., Manz, A., Nagel, S., Waltsgott, H. & Dwyer, J.H. (1995) Exposure to polychlorinated dioxins and furans (PCDD/F) and mortality in a cohort of workers from a herbicide producing plant in Hamburg, FRG. Am. J. Epidemiol., 142, 1165–1175.

Flesch-Janys, D., Berger, J., Konietzko, J., Gurn, P., Manz, A., Nagel, S., Waltsgott, H. & Dwyer, J.H. (1996) Erratum. Am. J. Epidemiol., 144, 716.

Flesch-Janys, D., Steindorf, K., Gurn, P. & Becher, H. (1998) Estimation of the cumulated exposure to polychlorinated dibenzo-p-dioxins/furans and standardized mortality ratio analysis of cancer mortality by dose in an occupationally exposed cohort. Environ. Health Perspectives, 106, 655–662.

Food Standards Agency (2000) Dioxins and PCBs in the UK diet: 1997 total diet study samples (Food Surveillance Information Sheet No. 4/00), London. Available at: <http:// www. foodstandards. gov. uk/ maff/ archive/ food/ infsheet>

Fox, T.R., Best, L.L., Goldsworthy, S.M., Mills, J. & Goldsworthy, T. (1993) Gene expression and cell proliferation in rat liver after 2,3,7,8-tetrachlorodibenzo-p-dioxin exposure. Cancer Res., 53, 2265–2271.

Fréry, N., Deloraine, A., Dor, F., Zeghoun, A. & Rouvière, F. (2000) Dioxins and furans in breast milk in France. Organohalogen Compd, 48, 47–50.

Fries, G.F. & Marrow, G.S. (1975) Retention and excretion of 2,3,7,8-tetrachlorodibenzo-p-dioxin by rats. J. Agric. Food Chem., 23, 265–273.

Gasiewicz, T.A., Geiger, L.E., Rucci, G. & Neal, R.A. (1983) Distribution, excretion and metabolism of 2,3,7,8-tetrachlorodibenzo-p-dioxin in C57BL/6J, DBA/2J and B6D2F1/j mice. Drug Metab. Disposition, 11, 397–403.

Gehrs, B.C. & Smialowicz, R.J. (1997) Alterations in the developing immune system of the F344 rat after perinatal exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin. I. Effects on the fetus and the neonate. Toxicology, 122, 219–228.

Gehrs, B.C. & Smialowicz, R.J. (1999) Persistent suppression of delayed-type hypersensitivity in adult F344 rats after perinatal exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin. Toxicology, 134, 79–88.

Gehrs, B.C., Riddle, M.M., Williams, W.C. & Smialowicz, R.J. (1997) Alterations in the developing immune system of the F344 rat after perinatal exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin. II. Effects on the pup and the adult. Toxicology, 122, 229–240.

Geiger, L.E. & Neal, R.A. (1981) Mutagenicity testing of 2,3,7,8-tetrachlorodibenzo-p-dioxin in histidine auxotrophs of Salmonella typhimurium. Toxicol. Appl. Pharmacol., 59, 125–129.

Gerschenson, L.E. & Rotello, R.J. (1992) Apoptosis: A different type of cell death. FASEB J., 6, 2450–2455.

Geusau, A., Tschachler, E., Meixner, M., Sandermann, S., Päpke, O., Wolf, C., Valic, E., Stingl, G. & McLachlan, M. (1999) Olestra increases faecal excretion of 2,3,7,8-tetrachlorodibenzo-p-dioxin. Lancet, 354, 1266–1267.

Geyer, H.J., Scheuntert, I., Rapp, K., Kettrup, A., Korte, F., Greim, H. & Rozman, K. (1990) Correlation between acute toxicity of TCDD and total body fat content in mammals. Toxicology, 65, 97–107.

Gochfeld, M., Nygren, M, Hansson, M., Rappe, C., Veléz, H., Ghent-Guenther, T., Wilson, W.P. & Kahn, P.C. (1989) Correlation of adipose and blood levels of several dioxin and dibenzofuran congeners in Agent Orange exposed Vietnam veterans. Chemosphere, 18, 517–524.

Gorski, T., Konopka, L. & Brodzki, M. (1984) Persistence of some polychlorinated dibenzo-p-dioxins and polychlorinated dibenzofurans of pentachlorophenol in human adipose tissue. Roczn. Pnstw. Zakl. Hyg., 35, 297–301.

Graham, M., Hileman, R., Orth, G., Wendling, J.M. & Wilson, J.D. (1986) Chlorocarbons in adipose tissue from a Missouri population. Chemosphere, 15, 1595–1600.

Gray, L.E., Jr & Ostby, J.S. (1995) In utero 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) alters reproductive morphology and function in female rat offspring. Toxicol. Appl. Pharmacol., 133, 285–294.

Gray, L.E., Jr, Kelce, W.R., Monosson, E., Ostby, J.S. & Birnbaum, L.S. (1995) Exposure to TCDD during development permanently alters reproductive function in male Long Evans rats and hamsters: Reduced ejaculated and epididymal sperm numbers and sex accessory gland weights in offspring with normal androgenic status. Toxicol. Appl. Pharmacol., 131, 108–118.

Gray, L.E., Jr, Ostby, J.S. & Kelce, W.R. (1997a) A dose–response analysis of the reproductive effects of a single gestational dose of 2,3,7,8-tetrachlorodibenzo-p-dioxin in male Long Evans hooded rat offspring. Toxicol. Appl. Pharmacol., 146, 11–20.

Gray, L.E., Jr, Wolf, C., Mann, P. & Ostby, J.S. (1997b) In utero exposure to low doses of 2,3,7,8-tetrachlorodibenzo-p-dioxin alters reproductive development of female Long Evans hooded rat offspring. Toxicol. Appl. Pharmacol., 146, 237–244.

Green, L.M. (1991) A cohort mortality study of forestry workers exposed to phenoxy acid herbicides. Br. J. Ind. Med., 48, 234–238.

Greig, J.B., Jones, G., Butler, W.H. et al. (1973) Toxic effects of 2,3,7,8-tetrachloro-dibenzo-p-dioxin. Food Cosmet. Toxicol., 11, 585–595.

Grubbs, W.D., Lustik, M.B., Brockman, A.S., Henderson, S.C., Burnett, F.R., Land, R.G., Osborn, D.J., Rocconi, V.K., Schrieber, M.E., Williams, D.E., Wolfe, W.H., Michalek, J.E., Miner, J.C., Henriksen, G.L. & Swaby, J.A. (1995) The Air Force Health Study: An epidemiologic investigation of health effects in Air Force personnel following exposure to herbicides. 1992 follow-up examination results (NTIS AD A-304-306–AD A-304-316), San Antonio: Armstrong Laboratory, Brooks Air Force Base.

Guo, Y.L., Lai, T.J., Ju, S.H., Chen, Y.C. & Hsu, C.C. (1993) Sexual developments and biological findings in Yucheng children. Organohalogen Compd, 14, 235–238.

Guo, Y.L., Lambert, G.H. & Hsu, C.C. (1995) Growth abnormalities in the population exposed in utero and early postnatally to polychlorinated biphenyls and dibenzofurans. Environ. Health Perspectives, 103 (Suppl. 6), 117–122.

Guo, Y.L., Hsu, P.C., Hsu, C.C. & Lambert, G.H. (2000) Semen quality after prenatal exposure to polychlorinated biphenyls and dibenzofurans. Lancet, 356, 1240–1241.

Haddad, S., Poulin, P. & Krishnan, K. (2000) Relative lipid content as the sole mechanistic determinant of the adipose tissue:blood partition coefficients of highly lipophilic organic chemicals. Chemosphere, 40, 839–843.

Häkansson, H., Johansson, L., Manzoor, E. & Ahlborg, U.G. (1991) Effects of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) on the vitamin A status of Hartley guinea pigs, Sprague-Dawley rats, C57Bl/2 mice, DBA/2 mice, and golden Syrian hamsters. J. Nutr. Sci. Vitaminol., 37, 117–138.

Halperin, W., Kalow, W., Sweeney, M.H., Tang, B.K., Fingerhut, M., Timpkins, B. & Wille, K. (1995) Induction of P-450 in workers exposed to dioxin. Occup. Environ. Med., 52, 86–91.

Hanberg, A., Häkansson, H. & Ahlborg, U.G. (1989) ‘ED50’ values for TCDD-induced reduction of body weight gain, liver enlargement, and thymic atrophy in Hartley guinea pigs, Sprague-Dawley rats, C57BL/6 mice, and golden Syrian hamsters. Chemosphere, 19, 813–816.

Hansson, M., Rodriguez-Pichardo, A., Smith, A.G., Greig, J.A. & Rappe, C. (1995) Levels of PCDDs and PCDFs in blood plasma from a Spanish family exposed to contaminated oil. Organohalogen Compd, 26, 205–208.

Harada, M. (1976) Intrauterine poisoning: Clinical and epidemiological studies of the problem. Bull. Inst. Constitut. Med., 25, 1–69.

Hardell, L., Eriksson, M., Lenner, P. & Lungren, E. (1981) Malignant lymphoma and exposure to chemicals, especially organic solvents, chlorophenols and phenoxy acids: A case–control study. Br. J. Cancer, 43, 169–176.

Harper, N., Connor, K., Steinberg, M. & Safe, S. (1994) An enzyme-linked immunosorbent assay (ELISA) specific for antibodies to TNP-LPS detects alterations in serum immunoglobulins and isotype switching in C57BL/6 and DBA/2 mice exposed to 2,3,7,8-tetrachlorodibenzo-p-dioxin and related compounds. Toxicology, 92, 155–167.

Hashimoto, S., Yamamoto, T., Yasuhara, A. & Morita, M. (1995) PCDD, PCDF, planar and other PCB levels in human milk in Japan. Chemosphere, 31, 4067–4075.

Hayabuchi, H., Yoshimura, T. & Kuratsune, M. (1979) Consumption of toxic rice oil by ‘Yusho’ patients and its relation to the clinical response and latent period. Food Cosmet. Toxicol., 17, 455–461.

Hecht, H. & Blüthgen, A. (1998) [Polychlorinated Dibenzo-p-dioxins and Furans in Meat and flseisherzuegnissen], Kulmbach: Institüt für chemie und physik der Bundesanstalt für Fleischforschung (in German).

Henck, J.M., New, M.A., Kociba, R.J. & Rao, K.S. (1981) 2,3,7,8-Tetrachlorodibenzo-p-dioxin: Acute oral toxicity in hamsters. Toxicol. Appl. Pharmacol., 59, 405–407.

Henriksen, G.L., Michalek, J.E., Swaby, J.A. & Rahe, A.J. (1996) Serum dioxin, testosterone, and gonadotropins in veterans of Operation Ranch Hand. Epidemiology, 7, 352–357.

Henriksen, G.L., Ketchum, N.S., Michalek, J.E. & Swaby, J.A. (1997) Serum dioxin and diabetes mellitus in veterans of Operation Ranch Hand. Epidemiology, 8, 252–258.

van den Heuvel, J.P., Clark, G.C., Kohn, M.C., Tritscher, A.M., Greenlee, W.F., Lucier, G.W. & Bell, D.A. (1994) Dioxin-responsive genes: Examination of dose–response relationships using quantitative reverse transcriptase-polymerase chain reaction. Cancer Res., 54, 62–68.

van den Heuvel, J.P., Clark, G.C., Tritscher, A.M. & Lucier, G.W. (1994) Accumulation of polychlorinated dibenzo-p-dioxins and dibenzofurans in liver of control laboratory rats. Fundam. Appl. Toxicol., 23, 465–469.

Hill, R.N., Erdreich, L.S., Paynter, O.E., Roberts, P.A., Rosenthal, S.L. & Wilkinson, C.F. (1989) Thyroid follicular cell carcinogenesis. Toxicol. Appl. Pharmacol., 12, 629–697.

Hilscherova, K., Machala, M., Kannan, K., Blankenship, A.L. & Giesy, J.P. (2000) Cell bioassays for detection of aryl hydrocarbon (AhR) and estrogen receptor (ER) mediated activity in environmental samples. Environ. Sci. Pollut. Res., 7, 159–171.

Hinsdill, R.D., Couch, D.L. & Speirs, R.S. (1980) Immunosuppression in mice induced by dioxin (TCDD) in feed. J. Environ. Pathol. Toxicol., 4, 401–425.

Hirakawa, H., Iida, T., Matsueda, T., Nakagawa, R., Hori, T. & Nagayama, J. (1995) Comparison of concentrations of PCDDs, PCDFs, PCBs and other organohalogen compound in human milk of primiparas and multiparas. Organohalogen Compd, 26, 197–200.

Holcomb, J., Ferrario, J. & Byrne, C. (1999) Analysis of mined clay products for PCDDs/PCDFs by high-resolution mass spectrometry and quadrupole ion storage mass spectrometry/mass spectrometry (QIMS/MS). Organohalogen Compd, 40, 137–140.

Holladay, S.D., Lindstrom, P., Blaylock, B.L., Comment, C.E., Germolec. D.R., Heindell, J.J. & Luster, M.I. (1991) Perinatal thymocyte antigen expression and postnatal immune development altered by gestational exposure to tetrachlorodibenzo-p-dioxin (TCDD). Teratology, 44, 385–393.

Holsapple, M.P. (1995) Immunotoxicity of halogenated aromatic hydrocarbons. In: Smialowicz, R.J. & Holsapple, M.P., eds, Experimental ImmunoToxicology, Boca Raton, FL: CRC Press, pp. 265–305.

Hoogenboom, L.A.P. (2001) Food chemical safety. In: Watson, D.H., ed., Contaminants, Weimar, Texas: Culinary and Hospitality Industry Publications Service, pp. 91–106.

Hoogenboom, L.A.P., Portier, L., van der Jagt, R.J. & Traag, W.A. (2000) Screening methods for dioxins. Bull. Int. Dairy Fed., 351, 19–23.

Hoogerbrugge, R. & Liem, A.K.D. (2000) Organohalogen Compd, 45, 13–16.

Hoogerbrugge, R., Freijer, J.I., Liem, A.K.D., Traag, W.A., Hoogenboom, L.A.P. & van Klaveren, J.D. (2000) Relationship between Critical Levels and Estimated Percentage of Non-compliance and Effect on Intakes (RIVM report 266/00 LOC RH), Bilthoven: National Institute of Public Health and the Environment.

Hooiveld, M., Heederik, D.J.J., Kogevinas, M., Boffetta, P., Needham, L.L., Patterson, D.G. & Bueno de Mesquita, H.B. (1998) Second follow-up of a Dutch cohort occupationally exposed to phenoxy herbicides, chlorophenols, and contaminants. Am. J. Epidemiol., 147, 891–901.

House, R.V., Lauer, L.D., Murray, M.J., Thomas, P.T., Ehrlich, J.P., Burleson, G.R. & Dean, J.H. (1990) Examination of immune parameters and host resistance mechanisms in B6C3F1 mice following adult exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin. J. Toxicol. Environ. Health, 31, 203–215.

Hsieh, S.-F., Yen, Y.-Y., Lan, S.-J., Hsieh, C.-C., Lee, C.-H. & Ko, Y.-C. (1996) A cohort study on mortality and exposure to polychlorinated biphenyls. Arch. Environ. Health, 51, 417–424.

Hsu, M.M.-L., Mak, C.-P. & Hsu, C.-C. (1995) Follow-up of skin manifestations in Yu-cheng children. Br. J. Dermatol., 132, 427–432.

Huisman, M., Koopman-Esseboom, C., Lanting, C.I., van der Paauw, C.G., Tuinstra, L.G., Fidler, V., Weisglas-Kuperus, N., Sauer, P.J., Boersma, E.R. & Touwen, B.C. (1995) Neurological condition in 18-month-old children perinatally exposed to polychlorinated biphenyls and dioxins. Early Hum. Dev., 43, 165–176.

Hurst, C., DeVito, M., Abbott, B. & Birnbaum L.S. (1997) Dose response of 2,3,7,8- tetrachlorodibenzo-p-dioxin (TCDD) distribution in pregnant Long Evans rats. Toxicologist, 36, 257.

Hurst, C.H., Abbott, B.D., Devito, M.J. & Birnbaum, L.S. (1998a) 2,3,7,8-Tetrachlorodibenzo-p-dioxin in pregnant Long Evans rats: Disposition to maternal and embryo/fetal tissues. Toxicol. Sci., 45, 129–136.

Hurst, C.H., DeVito, M.J. & Birnbaum, L.S. (1998b) Tissue disposition of 2,3,7,8- tetrachlorodi-benzo-p-dioxin (TCDD) in maternal and developing Long Evans rats following subchronic exposures. Toxicologist, 42, 258.

Hurst, C.H., De Vito, M.J., Setzer, R.W. & Birnbaum, L. (2000a) 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.

Hurst, C.H., DeVito, M.J. & Birnbaum, L.S. (2000b) Tissue disposition of 2,3,7,8-tetrachlorodi-benzo-p-dioxin (TCDD) in maternal and developing Long-Evans rats following subchronic exposure. Toxicol. Sci., 57, 275–283.

IARC (1987) IARC Monographs on the Evaluation of the Carcinogenic Risk of Chemicals to Humans, Suppl. 7, Overall evaluations of carcinogenicity: An updating of IARC Monographs Volumes 1–29, Lyon: IARCPress.

IARC (1997) IARC Monographs on the Evaluation of the Carcinogenic Risk of Chemicals to Humans, Vol. 69, Polychlorinated dibenzo-para-dioxins and polychlorinated dibenzofurans, Lyon: IARCPress.

Ideo, G., Ballati, G., Bellobuno, A. & Bissanti, L. (1985) Urinary D-glucaric excretion in the Seveso area, polluted by tetrachlorodibenzo-p-dioxin (TCDD): Five years of experience. Environ. Health Perspectives, 60, 151–157.

Iida, T., Hirakawa, H., Matsueda, T., Nagayama, J. & Nagata, T. (1999a) Polychlorinated dibenzo-p-dioxins and related compounds: Correlations of levels in human tissues and in blood. Chemosphere, 38, 2767–2774

Iida, T., Hirakawa, H., Matsueda, T., Takenaka, S. & Nagayama, J. (1999b) Polychlorinated dibenzo-p-dioxins and related compounds: The blood levels in young Japanese women. Chemosphere, 38, 3497–3502

Ilsen, A., Briet, J.M., Koppe, J.G., Pluim, H.J. & Oosting, J. (1996) Signs of enhanced neuromotor maturation in children due to perinatal load with background levels of dioxins. Follow-up until age 2 years and 7 months. Chemosphere, 33, 1317–1326.

Ishida, C., Koga, N, Saeki, H.K. & Yoshimura, H. (1991) Metabolism of 3,4,3’,4’- and 2,5,2’,5’-tetrachlorobiphenyl by rat microsomes and highly purified cytochrome P450. J. Pharmacobio-Dynamics, 14, 276–284.

Jan, J. & Vrbi, V. (2000) Polychlorinated biphenyls cause developmental enamel defects in children. Caries Res., 34, 469–473.

Jobst, H. & Aldag, R. (2000) Dioxine in Lagerstätten-Tonen. Z. Umweltchem., 12, 2–4.

Johnson, E.S. (1990) Association between soft tissue sarcomas, malignant lymphomas, and phenoxy herbicides/chlorophenols: Evidence from occupational cohort studies. Fundam. Appl. Toxicol., 14, 219–234.

Johnson, E.S., Parsons, W., Weinberg, C.R., Shore, D.L., Mathews, J., Patterson, D.K., Jr & Needham, L.L. (1992) Current serum levels of 2,3,7,8-tetrachlorodibenzo-p-dioxin in phenoxy acid herbicide applicators and characterization of historical levels. J. Natl Cancer Inst., 84, 1648–1653.

Johnson, K.L., Cummings, A.M. & Birnbaum, L.S. (1997) Promotion of endometriosis in mice by polychlorinated dibenzo-p-dioxins, dibenzofurans, and biphenyls. Environ. Health Perspectives, 105, 750–755.

Jones, K.G. & Sweeney, G.D. (1980) Dependence of the porphyrinogenic effect of 2,3,7,8-tetrachlorodibenzo-p-dioxin upon inheritance of aryl hydrocarbon hydroxylase responsiveness. Toxicol. Appl. Pharmacol., 53, 42–49.

Jurek, M.A., Powers, R.H., Gilbert, L.G. & Aust, S.D. (1990) The effect of TCDD on acyl CoA:retinol acyltransferase activity and vitamin A accumulation in the kidney of male Sprague-Dawley rats. J. Biochem. Toxicol., 5, 155–160.

Kanimura, H., Koga, N., Oguri, K., Yoshimura, H., Honda, Y. & Nakano, M. (1988) Enhanced faecal excretion of 2,3,4,7,8-pentachlorodibenzofuran in rats by a long-term treatment with activated charcoal beads. Xenobiotica, 18, 585–574.

Kang, Y.S., Matsuda, M., Kawano, M., Wakimoto, T. & Min, B.Y. (1997) Organochlorine pesticides, polychlorinated biphenyls, polychlorinated dibenzo-p-dioxins and dibenzofurans in human adipose tissue from western Kyungnam, Korea. Chemosphere, 35, 2107–2117.

Kashani, M., Steiner, G., Haitel, A., Schaufler, K., Thalhammer, T., Amann, G., Kramer, G., Marberger, M. & Scholler, A. (1998) Expression of the aryl hydrocarbon receptor (AhR) and the aryl hydrocarbon receptor nuclear translocator (ARNT) in fetal, benign hyperplastic, and malignant prostate. Prostate, 37, 98–108.

Kedderis, L.B., Diliberto, J.J., Linko, P., Goldstein, J.A. & Birnbaum, L.S. (1991) Disposition of 2,3,7,8-tetrabromodibenzo-p-dioxin and 2,3,7,8-tetrachlorodibenzo-p-dioxin in the rat: Biliary excretion and induction of cytochrome CYP1A1 and CYP1A2. Toxicol. Appl. Pharmacol., 111, 163–172.

Kelley, S.K., Nilsson, C.B., Green, M.H., Green, J.B. & Hakansson, H. (2000) Mobilization of vitamin A stores in rats after administration of 2,3,7,8-tetrachlorodibenzo-p-dioxin: A kinetic analysis. Toxicol. Sci., 55, 478–484.

Kerkvliet, N.I. & Brauner, J.A. (1987) Mechanisms of 1,2,3,4,6,7,8-heptachlorodibenzo-p-dioxin (HpCDD)-induced humoral immune suppression: Evidence of primary defect in T cell regulation. Toxicol. Appl. Pharmacol., 87, 18–31.

Kerkvliet, N.I. & Burleson, G.R. (1994) Immunotoxicity of TCDD and related halogenated aromatic hydrocarbons. In: Dean, J.H., Luster, M.I., Munson, A.E. & Kimber, I., eds, ImmunoToxicology and Immunopharmacology, 2nd Ed., New York: Raven Press, pp. 97–121.

Ketchum, N.S., Michalek, J.E. & Burton, J.E. (1999) Serum dioxin and cancer in veterans of Operation Ranch Hand. Am. J. Epidemiol., 149, 630–639.

King, F.G., Dedrick, R.L., Collins, J.M., Matthews, H.B. & Birnbaum, L.S. (1983) Physiological model for the pharmacokinetics of 2,3,7,8-tetrachlorodibenzofuran in several species. Toxicol. Appl. Pharmacol., 67, 390–400.

Kiviranta, H., Vartiainen, T., Verta, M., Tuomisto, J.T. & Tuomisto, J. (2000) High fish-specific dioxin concentrations in Finland. Lancet, 27, 1883–1885.

Kociba, R.J., Keeler, P.A., Park, C.N. & Gehring, P.J. (1976) 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD): Results of a 13-week oral toxicity study in rats. Toxicol. Appl. Pharmacol., 35, 553–574.

Kociba, R.J., Keyes, D.G., Beyer, J.E. & Carreon, R.M. (1978a) Toxicologic studies of 2,3,7,8-tetrachloro-p-dioxin (TCDD) in rats. Toxicol. Occup. Med., 4, 282–287.

Kociba, R.J., Keyes, D.G., Beyer, J.E., Carreon, R.M., Wade, Dittenberg, D.A., Kalnis, R.P., Franson, L.E., Park, C.N., Bernard, S.D., Hummel, R.A. & Humiston, C.G. (1978b) Results of a two-year chronic toxicity and oncogenicity study of 2,3,7,8-tetrachlorodibenzo-p-dioxin in rats. Toxicol. Appl. Pharmacol., 46, 279–303.

Kogevinas, M. (2000) Studies of cancer in humans. Food Addit. Contam., 17, 317–324.

Kogevinas, M., Kauppinen, T., Winkelmann, R., Becher, H., Bertazzi, P.A., Bueno de Mesquita, H.B., Coggon, D., Green, L.M., Johnson, E.S., Littorin, M., Lynge, E., Marlow, D.A., Mathews, J.D., Neuberger, M., Benn, T., Pannett, P., Pearce, N.E. & Saracci, R.R. (1995) Soft tissue sarcoma and non-Hodgkin’s lymphoma in workers exposed to chlorophenoxy herbicides, chlorophenols and dioxins. Two nested case–control studies. Epidemiology, 6, 396–402.

Kogevinas, M., Becher, H., Benn, T., Bertazzi, P.A., Boffetta, P., Bueno-de-Mesquita, H.B., Coggon, D., Colin, D., Flesch-Janys, D., Fingerhut, M., Green, L., Kauppinen, T., Littorin, M., Lynge, E., Mathews, J.D., Neuberger, M., Pearce, N. & Saracci, R. (1997) Cancer mortality in workers exposed to phenoxy herbicides, chlorophenols and dioxins. An expanded and updated international cohort study. Am. J. Epidemiol., 145, 1061–1075.

Kohn, M.C., Lucier, G.W., Clark, G.C., Sewall, C., Tritscher, A.M. & Portier, C.J. (1993) A mechanistic model of effects of dioxin on gene expression in the rat liver. Toxicol. Appl. Pharmacol., 120, 138–154.

Kohn, M.C., Lucier, G.W. & Portier, C.J. (1994) The importance of biological realism in dioxin risk assessment models. Risk Anal., 14, 993–1000.

Kohn, M.C. Sewall, C.H., Lucier, G.W. & Portier, C.J. (1996) A mechanistic model of effects of dioxin on thyroid hormones in the rat. Toxicol. Appl. Pharmacol., 165, 29–48.

Koopman-Esseboom, C., Morse, D.C., Weisglas-Kuperus, N., Lutkeschipholt, I.J., van der Paauw, C.G., Tuinstra, L.G., Brouwer, A. & Sauer, P.J. (1994) Effects of dioxins and polychlorinated biphenyls on thyroid hormone status of pregnant women and their infants, Pediatr. Res., 36, 468–473.

Koopman-Esseboom, C., Weisglas-Kuperus, N., de Ridder, M.A.J., van der Paauw, C.G., Tuinstra, L.G.M.T. & Sauer, P.J.J. (1996) Effects of polychlorinated biphenyl/dioxin exposure and feeding type on infants’ mental and psychomotor development. Pediatrics, 97, 700–706.

Körner, W. & Hagenmaier, H. (1991) Report for the Ministry of the Environment, Baden-Württemberg.

Koshakji, R.P., Harbison, R.D. & Bush, M.T. (1984) Studies on the metabolic fate of 14C-2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in the mouse. Toxicol. Appl. Pharmacol., 73, 69–77.

Koss, G., Meyer-Rogge, D., Seubert, S., Seubert, A. & Losekam, M. (1993) 2,2’,3’,4,4’,5,5’-Heptachlorobiphenyl (PCB 180). On its toxicokinetics, biotransformation and porphyrinogenic action in female rats. Arch. Toxicol., 67, 651–654.

Kozak, K.R., Abbott, B. & Hankinson, O. (1997) ARNT-deficient mice and placental differentiation. Dev. Biol., 191, 297–305.

Kreuzer, P.E., Csanády, G.A., Bauer, C., Kessler, W., Päpke, O., Greim, H. & Filser, J.G. (1997) 2,3,7,8-Tetrachlorodibenzo-p-dioxin (TCDD) and congeners in infants. A toxicokinetic model of human lifetime body burden by TCDD with special emphasis on its uptake by nutrition. Arch. Toxicol., 71, 383–400.

Kunita, N., Hori, S., Obana, H., Otake, T., Nishimura, H., Kashimoto, T. & Ikegami, N. (1985) Biological effect of PCBs, PCQs, and PCDFs present in the oil causing Yusho and Yu-cheng. Environ. Health Perspectives, 59, 79–84.

Kuratsune, M., Yoshimura, T., Matsuzaka, J. & Yamaguchi, A. (1972) Epidemiologic study on Yusho—A poisoning caused by ingestion of rice oil contaminated with a commercial brand of polychlorinated biphenyls. Environ. Health Perspectives, 46, 119–128.

Kuratsume, M., Yoshimura, H., Hori, Y., Okumura, M. & Masuda, Y. (1996) Yusho: A Human Disaster Caused by PCBs and related Compounds, Fukuoka, Kyushu University Press.

Lai, T.J. et al. (2001) Effect of prenatal exposure to polychlorinated biphenyls on cognitive development in children: A longitudinal study in Taiwan. Br. J. Psychiatr., 178 (Suppl. 40), s49–s52.

Lakshmanan, M.R., Campbell, B.S., Chirtel, S.J., Ekarohita, N. & Ezekiel, M. (1986) Studies on the mechanism of absorption and distribution of 2,3,7,8-tetrachlorodibenzo-p-dioxin in the rat. J. Pharmacol. Exp. Ther., 239, 673–677.

Lan, S.J., Yen, Y.Y., Lan, J.L., Chen, E.R. & Ko, Y.C. (1990) Immunity of PCB transplacental Yu-cheng children in Taiwan. Bull. Environ. Contam. Toxicol., 44, 224–229.

Landi, M.T., Needham, L.L., Lucier, G., Mocarelli, P., Bertazzi, P.A. & Caporaso, N. (1997) Concentrations of dioxin 20 years after Seveso. Lancet, 349, 1811.

Lans, M.C., Spiertz, C., Brouwer, A. & Koeman, J.H. (1994) Different competition of thyroxine binding to thransthyretin and thyroxine-binding globulin by hydroxy-PCBs, PCDDs and PCDFs. Eur. J. Pharmacol. Environ. Toxicol. Pharmacol., 270, 129–136.

LeBel, G.L., Williams, D.T., Benoit, F.M. & Goddard, M. (1990) Polychlorinated dibenzodioxins and dibenzofurans in human adipose tissue samples from five Ontario municipalities. Chemosphere, 21, 1465–1475.

van Leeuwen, F.X.R. & Younes, M.M., eds (2000) Assessment of the health risk of dioxins: Re-evaluation of the tolerable daily intake (TDI). Food Addit. Contam., 17.

Leung, H.W., Ku, R.H., Paustenbach, D.J. & Andersen, M.E. (1988) A physiologically based pharmacokinetic model for 2,3,7,8-tetrachlorodibenzo-p-dioxin in C57BL/6J and DBA/2J mice. Toxicology Lett., 42, 15–28.

Leung, H.W., Paustenbach, D.J., Murray, F.J. & Andersen, M.E. (1990a) A physiological pharmacokinetic description of the tissue distribution and enzyme inducing properties of 2,3,7,8-tetrachlorodibenzo-p-dioxin in the rat. Toxicol. Appl. Pharmacol., 103, 399–410.

Leung, H.W., Poland, A., Paustenbach, D.J., Murray, F.J. & Andersen, M.E. (1990b) Pharmacokinetics of [125I]-2-iodo-3,7,8-trichlorodibenzo-p-dioxin in mice: Analysis with a physiologically modelling approach. Toxicol. Appl. Pharmacol., 103, 411–419

Leung, H.W., Wendling, J.M., Orth, R., Hileman, F. & Paustenbach, J.D. (1990c) Relative distribution of 2,3,7,8-tetrachlorodibenzo-p-dioxin in human and human tissues. Toxicol. Lett., 50, 275–282.

Li, X. & Rozman, K.K. (1995) Subchronic effects of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) and their reversibility in male Sprague-Dawley rats. Toxicology, 97, 133–140.

Li, X., Johnson, D.C. & Rozman, K.K. (1995) Reproductive effects of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in female rats: Ovulation, hormonal regulation, and possible mechanism(s). Toxicol. Appl. Pharmacol., 133, 321–327.

Li, W., Wu, W.Z., Schramm, K.-W., Xu, Y. & Kettrup, A. (1999) Toxicity of mixtures of polychlorinated dibenzo-p-dioxins, dibenzofurans, and biphenyls determined by dose–response curve analysis. Bull. Environ. Contam. Toxicol., 62, 539–546.

Liem, A.K.D., Theelen, R.M.C. & Hoogerbrugge, R.(1996) Dioxinen en PCB’s in voeding. Resultaten van aanvullend onderzoek (RIVM report 639102.005), Bilthoven: National Institute of Public Health and the Environment (in Dutch with English summary).

Liem, A.K.D., Hoogerbrugge, R., Cuijpers, C.E.J., den Hartog, R.S., Hijman, W.C., Linders, H.M.A., Marsman, J.A., van der Velde, E.G. & Zomer, B. (1997) Trends in dietary exposure and PCBs in The Netherlands. Results from a duplicate diet study. Organohalogen Compd, 33, 112–115.

Lindström, G., Haug, L.S. & Nicolaysen, T. (2000) Intercalibration on Dioxin in Food, Oslo: National Institute of Public Health.

Lipp, H.-P., Schrenk, D., Wiesmüller, T., Hagenmaier, H. & Bock, K.W. (1992) Assessment of biological activities of mixtures of polychlorinated dibenzo-p-dioxins (PCDDs) and their constituents in human HEPG2 cells. Arch. Toxicol., 66, 220–223.

Littorin, M., Hansson, M., Rappe, C. & Kogevinas, M. (1994) Dioxins in blood from Swedish phenoxy herbicide workers. Lancet, 344, 611–612.

Lu, Y.-C. & Wu, Y.-C. (1985) Clinical findings and immunological abnormalities in Yu-cheng patients. Environ. Health Perspectives, 59, 17–29.

Lucier, G.W. (1991) Humans are a sensitive species to some of the biochemical effects of structural analogs of dioxin. Environ. Toxicol., 10, 727–735.

Lucier, G.W., Tritscher, A., Goldsworthy, T., Foley, J., Clark, G. & Goldstein, J. (1991) Ovarian hormones enhance 2,3,7,8-tetrachlorodibenzo-p-dioxin mediated increases in cell proliferation and preneoplastic foci in a two-stage model for rat hepatocytes. Cancer Res., 51, 1391–1397.

Luebke, R.W., Copeland, C.B. & Andrews, D.L. (1995) Host resistance to T. spiralis infection in rats exposed to 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD). Fundam. Appl. Toxicol., 24, 285–289.

Luebke, R.W., Copeland, C.B., Diliberto, J.J., Akubue, P.I., Andrews, D.L., Riddle, M.M., Williams, W.C. & Birnbaum, L.S. (1994) Assessment of host resistance to Trichinella spiralis in mice following pre-infection exposure to 2,3,7,8-TCDD. Toxicol. Appl. Pharmacol., 125, 7–16.

Lundgren, K., Andries, M., Thompson, C. & Lucier, G.W. (1986) Dioxin treatment of rats results in increased in vitro induction of sister chromatid exchanges by alpha-naphthoflavone: An animal model for human exposure to halogenated aromatics. Toxicol. Appl. Pharmacol., 85, 189–195.

Luster, M.I., Boorman, G.A., Harris, M.W. & Moore, J.A. (1980) Laboratory studies on polybrominated biphenyl-induced immune alterations following low-level chronic and pre-/postnatal exposure. Int. J. Immunopharmacol., 2, 69–80.

Lynge, E. (1993) Cancer in phenoxy herbicide manufacturing workers in Denmark, 1947–87—An update. Cancer Causes Control, 4, 261–272.

Lynge, E. (1985) A follow-up study of cancer incidence among workers in manufacture of phenoxy herbicides in Denmark. Br. J. Cancer, 52, 259–270.

Mably, T.A., Moore, R.W. & Peterson, R.E. (1992a) In utero and lactational exposure of male rats to 2,3,7,8-tetrachlorodibenzo-p-dioxin. 1. Effects on androgenic status. Toxicol. Appl. Pharmacol., 114, 97–107.

Mably, T.A., Bjerke, D.L., Moore, RW., Gendron-Fitzpatrick, A. & Peterson, R.E. (1992b) In utero and lactational exposure of male rats to 2,3,7,8-tetrachlorodibenzo-p-dioxin. 3. Effects on spermatogenesis and reproductive capability. Toxicol. Appl. Pharmacol., 114, 118–126.

Mably, T.A., Moore, R.W., Goy, R.W. & Peterson, R.E. (1992c) In utero and lactational exposure of male rats to 2,3,7,8-tetrachlorodibenzo-p-dioxin. 2. Effects on sexual behavior and the regulation of luteinizing hormone secretion in adulthood. Toxicol. Appl. Pharmacol., 114, 108–117.

Malisch, R. (2000a) PCDD/F in kaolinitic clays and its relevance for feedingstuff, food and cosmetics. Organohalogen Compd, 47, 326–329.

Malisch, R. (2000b) Increase of the PCDD/F-contamination of milk, butter and meat samples by use of contaminated citrus pulp. Chemosphere, 40, 1041–1053.

Malisch, R., Schmid, P., Frommberger, R. & Fürst, P. (1996) Results of a quality control study of different analytical methods for determination of PCDD/PCDF in egg samples. Chemosphere, 32, 31–44.

Malisch, R., Bruns-Weller, E., Knoll, A., Thoma, H. & Peichl. L. (1997) Results of a quality control study of different analytical methods for determination of PCDD/PCDF in kale samples. Organohalogen Compd, 31, 83–88.

Malisch, R., Berger, B. & Verstraete, F. (1999) Lime as source for PCDD/F-contamination of Brazilian citrus pulp pellets (CPPs). Organohalogen Compd, 41, 51–53.

Malisch, R., Bruns-Weller, E., Knoll, A., Fürst, P., Mayer, R. & Wiesmüller, T. (2000) Results of an ‘emergency quality control study’ as confirmation of a PCDD/PCDF-contamination of milk and butter samples. Chemosphere, 40, 1033–1040.

Malisch, R., Baumann, B., Behnisch, P.A., Canady, R., Fraisse, P., Fürst, P., Hayward, D., Hoogenboom, R., Hoogerbrugge, R., Liem, D., Päpke, O., Traag, W. & Wiesmüller, T. (2001) Harmonized quality criteria for chemical and bioassay analyses of PCDDs/PCDFs in feed and food. Part 1: General considerations, GC/MS methods. Organohalogen Compd (in press).

Manz, A., Berger, J., Dwyer, J.H., Flesch-Janys, D., Nagel, S. & Waltsgott, H. (1991) Cancer mortality among workers in chemical plant contaminated with dioxin. Lancet, 338, 959–964.

Massa, T., Ismaeili, A., Fortmeyer, H. Schlatterer, B., Hagenmaier, H. & Chandra, P. (1992) Cell transforming and oncogenic activity of 2,3,7,8-tetrachloro- and 2,3,7,8-tetrabromodibenzo-p-dioxin. Anticancer Res., 12, 2053–2060.

Mastroiacovo, P., Spagnolo, A., Marni, E., Meazza, L., Bertollini, R., Segni, G. & Borgna-Pignatti, C. (1988) Birth defects in the Seveso area after TCDD contamination. J. Am. Med. Assoc., 259, 1668–1672.

Masuda, Y. (1992) Approach to risk assessment of PCDDs and PCDFs in Yusho food poisoning. Toxic Subst. J., 12, 175–180.

Masuda, Y., Kuroki, H., Haraguchi, K. & Nagayama, J. (1985) PCB and PCDF congeners in the blood and tissues of Yusho and Yu-cheng patients. Environ. Health Perspectives, 59, 53–58.

Mayani, A., Barel, S., Soback, S. & Almagor, M. (1997) Dioxin concentrations in women with endometriosis. Hum. Reprod., 12, 373–375.

Mayer, R. (1998) Organohalogen Compd, 38, 139–142.

Mayer, R. & Jahr, D. (1998) Lebensmittelchemie, 52, 100–104.

Mayura, K., Spainhour, C.B., Howie, L., Safe, S. & Phillips, T.D. (1993) Teratogenicity and immunotoxicity of 3,3',4,4',5-pentachlorobiphenyl in C57BL/6 mice. Toxicology; 77, 123–131.

McClain, R.M. (1989) The significance of hepatic microsomal enzyme induction and altered thyroid function in rats: Implications for thyroid gland neoplasia. Toxicol. Pathol., 17, 294–306.

McClain, R.M., Levin, A.A., Posch, R. & Downing, J.C. (1989) The effect of phenobarbital on the metabolism and excretion of thyroxine in rats. Toxicol. Appl. Pharmacol., 99, 216.

McConkey, D.J., Hartzell, P., Duddy, S.K., Håkansson, H. & Orrehius, S. (1988) 2,3,7,8-Tetrachlorodibenzo-p-dioxin kills immature thymocytes by Ca2+-mediated endonuclease activation. Science, 242, 256–259.

McConnell, E.E., Moore, J.A., Haseman, J.K. & Harris, M.W. (1978) The comparative toxicity of chlorinated dibenzo-p-dioxins in mice and guinea pigs. Toxicol. Appl. Pharmacol., 44, 335–356.

McGregor, D.B., Brown, A.G., Howgate, S., McBride, D., Riach, C. & Caspary, W.J. (1991) Response of the L5178Y mouse lymphoma cell forward mutation assay: V. 27 coded chemicals. Environ. Mol. Mutag., 17, 196–219.

McKinley, M.K., Kedderris, L.B. & Birnbaum, L.S. (1993) The effect of pretreatment on the biliary excretion of 2,3,7,8-tetrachloro-p-dioxin, 2,3,7,8-tetrachlorodibenzofuran and 3,3’,4,4’-tetrachlorobiphenyl in the rat. Fundam. Appl. Toxicol., 21, 425–432.

McLachlan, M.S. (1993) Digestive tract absorpion of polychlorinated dibenzo-p-dioxins and dibenzo-p-furans, and biphenyls in a nursing infant. Toxicol. Appl. Pharmacol., 123, 68–72.

McLachlan, M.S. & Riechter, W. (1998) Uptake and transfer of PCDD/Fs by cattle fed naturally contaminated feedstuffs and feed contaminated as a result of sewage sludge application. I. Lactating cows. J. Agric. Food Chem., 46, 1166–1172.

McNulty, W.P., Pomerantz, I. & Farrell, T. (1981) Chronic toxicity of 2,3,7,8-tetrachlorodibenzo-furan for rhesus macaques. Food Chem. Toxicol., 19, 57–65.

Meyne, J., Allison, D.C., Bose, K., Jordan, S.W., Rodolpho, P.F. & Smith, J. (1985) Hepatotoxic doses of dioxin do not damage mouse bone marrow chromosomes. Mutat. Res., 157, 63–69.

Michalek, J.E., Wolfe, W.H. & Miner, J.C. (1990) Health status of Air Force veterans occupationally exposed to herbicides in Vietnam. II. Mortality. J. Am. Med. Assoc., 264, 1832–1836.

Michalek, J.E., Pirkle, J.L., Caudill, S.P., Tripathi, R.C., Patterson, D.G. & Needham, L.L. (1996) Pharmacokinetics of TCDD in veterans of operation Ranch Hand: 10-year follow up. J. Toxicol. Environ. Health, 47, 209–220.

Michalek, J.E., Rahe, A.J. & Boyle, C.A. (1998a) Paternal dioxin, preterm birth, intrauterine growth retardation, and infant death. Epidemiology, 9, 161–167.

Michalek, J.E., Rahe, A. & Boyle, C. (1998b) Paternal dioxin and the sex of children fathered by veterans of operation Ranch Hand. Epidemiology, 9, 474–475.

Michalek, J.E., Ketchum, N.S. & Akhtar, F.Z. (1998c) Postservice mortality of US Air Force veterans occupationally exposed to herbicides in Vietnam: 15-year follow-up. Am. J. Epidemiol., 148, 786–792.

Ministry of Agriculture, Fisheries and Food (1997) Food Surveillance Information Sheet Number 105, London.

Ministry of Agriculture, Fisheries and Food (2000) National Food Survey 1999, London. Available at: <http://www.maff.gov.uk/esg/work_htm/publications/cf/nfs/nfs.htm>.

Miyata, H., Fukushima, S., Kashimoto, T. & Kunita, N. (1985) PCBs, PCQs and PCDFs in tissues of Yusho and Yu-cheng patients. Environ. Health Perspectives, 59, 67–72.

Mocarelli, P., Marocchi, A., Brambilla, P., Gerthoux, P.M., Young, D.S. & Mantel, N. (1986) Clinical laboratory manifestations of exposure to dioxin in children. A six year study of the effects of an environmental disaster near Seveso, Italy. J. Am. Med. Assoc., 256, 2687–2695.

Mocarelli, P., Brambilla, P., Gerthoux, P.M., Patterson, D.G., Jr & Needham, L.L. (1996) Change in sex ratio with exposure to dioxin. Lancet, 384, 409.

Mocarelli, P., Gerthoux, P.M., Ferrari, E., Patterson, D.G., Jr, Kieszak, S., Brambilla, P., Vincoli, N., Signorini, S., Tramacere, P., Carreri, V., Sampson, E.J., Turner, W.E. & Needham, L.L. (2000) Paternal concentrations of dioxin and sex ratio of offspring. Lancet, 355, 1858–1863.

van der Molen, G.W. (1998) A Physiologically-based Mathematical Model for the Long-term Kinetics of Dioxins and Furans in Humans, PhD Thesis, Free University, Amsterdam.

van der Molen, G.W., Kooijman, B.A.L.M., Wittspepe, J., Schrey, P., Flesch-Janys, D. & Slob, W. (2000) Estimation of dioxin and furan elimination rates with a pharmacokinetic model. J. Exposure Anal. Environ. Epidemiol., 10, 579–585.

Morita M., et al. (1977) Detailed examination of polychlorinated dibenzofurans in PCB preparations and Kanemi Yusho oil. Bull. Environ. Contam. Toxicol., 18, 67–73.

Mortelmans, K., Haworth, S., Speck, W. & Zeiger, E. (1984) Mutagenicity testing of Agent Orange components and related compounds. Toxicol. Appl. Pharmacol., 75, 137–146.

Murray, F.J., Smith, F.A., Nitschke, K.D., Humiston, C.G., Kociba, R.J. & Schwetz, B.A. (1979) Three-generation reproduction study of rats given 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in the diet. Toxicol. Appl. Pharmacol., 50, 241–252.

Nagayama, J., Nagayama, M. & Masuda, Y. (1993) Frequency of micronuclei induced in cultured lymphocytes by highly toxic organochlorine congeners. Fukuoka Acta Med., 84, 189–194.

Nagayama, J., Nagayama, M., Iida, T., Hirakawa, H., Matsueda, T. & Masuda, Y. (1994) Effects of highly toxic organochlorine compounds retained in human body on induction of sister chromatid exchanges in cultured human lymphocytes. Chemosphere, 29, 2349–2354.

Nagel, S., Berger, J., Flesch-Janys, D., Manz, A. & Ollroge, I. (1994) Mortality and cancer mortality in a cohort of female workers of a herbicide producing plant exposed to polychlorinated dibenzo-p-dioxins and furans. Inf. Biomet. Epidemiol. Med. Biol., 25, 32–38.

Narasimhan, T.R., Craig, A., Arellano, L., Harper, N., Howie, L., Menache, M., Birnbaum, L. & Safe, S. (1994) Relative sensitivities of 2,3,7,8-tetrachlorodibenzo-p-dioxin-induced Cyp1a-1 and Cyp1a-2 gene expression and immunotoxicity in female B6C3F1 mice. Fundam. Appl. Toxicol., 23, 598–607.

NATO (North Atlantic Treaty Organization) (1988) International toxicity equivalency factors (I-TEF) method of risk assessments of complex mixtures of dioxins and related compounds (Report No. 176). Brussels: Committee on the Challenges of Modern Society.

Needham, L.L., Gerthoux, P.M., Patterson, D.G., Brambilla, P., Pirkle, J.L., Tramacere, P.I., Turner, W.E., Beretta, C., Sampson, E.J. & Mocarelli, P. (1994) Half-life of 2,3,7,8-tetrachlorodibenzo-p-dioxin in serum of Seveso adults: Interim report. Organohalogen Compd., 21, 81–85.

Neuberger, M., Landvoigt, W. & Denrtl, F. (1991) Blood levels of 2,3,7,8-tetrachlorodibenzo-o-dioxin in chemical workers after chloracne and in comparison groups. Int. Arch. Occup. Environ. Health, 63, 325–327.

Neubert, D., Zens, P., Rothenwallner, A. & Merker, H.-J. (1973) A survey of the embryotoxic effects of TCDD in mammalian species. Environ. Health Perspectives, 5, 67–79.

Newsome, W.H. & Ryan, J.J. (1999) Toxaphene and other chlorinated compounds in human milk from northern and southern Canada: A comparison. Chemosphere, 39, 519–526.

Nilsson, C.B., Hanberg, A., Trossvik, C. & Hakansson, H. (1996) 2,3,7,8-Tetrachlorodi-benzo-p-dioxin affects retinol esterification in rat hepatic stellate cells and kidney. Environ. Toxicol. Pharmacol., 2, 17–23.

Nohara, K., Fujimaki, H., Tsukumo, S., Ushio, H., Miyabara, Y., Kijama, M., Tohyama, C. & Yonemoto, J. (2000) The effects of perinatal exposure to low doses of 2,3,7,8-tetrachlorodibenzo-p-dioxin on immune organs in rats. Toxicology, 154, 123–133.

Nygren, M., Rappe, C., Lindström, G., Hansson, M., Bergqvist, P.A., Marklund, S., Domellöf, L., Hardell, L. & Olsson, M. (1986) Identification of 2,3,7,8-substituted polychlorinated dioxins and dibenzofurans in environmental and human samples. In: Rappe, C., Choudhary, G. & Keith, L.H., eds, Chlorinated Dioxins and Dibenzofurans in Perspective, Chelsea, MI: Lewis Publishers, pp. 17–34.

Ohsako, S., Miyabara, Y., Nishimura, N., Kurosawa, S., Sakaue, M., Ishimura, R., Sato, M., Takeda, K., Aoki, Y., Sone, H., Tohyama, C. & Yonemoto, J. (2001) Maternal exposure to a low dose of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) suppressed the development of reproductive organs of male rats: Dose-dependent increase of mRNA levels of 5-reductase type 2 in contrast to decrease of androgen receptor in the pubertal ventral prostate. Toxicol. Sci., 60, 132–143.

Olson, J.R., Holscher, M.A. & Neal, R.A. (1980) Toxicity of 2,3,7,8-tetrachlorodibenzo-p-dioxin in the golden Syrian hamster. Toxicol. Appl. Pharmacol., 55, 67–78.

Ostby, J., Price, M., Huey, O., Hurst, C., Birnbaum, L. & Gray, L.E., Jr (1999) Developmental and reproductive effects of low-dose, steady-state maternal 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) administration. Toxicologist, 48, 147–152.

Ott, M.G. & Zober, A. (1996) Cause specific mortality and cancer incidence among employees exposed to 2,3,7,8-TCDD after a 1953 reactor accident. Occup. Environ. Med., 53, 606–612.

Ott, M.G., Olson, R.A., Cook, R.R. & Bond, G.G. (1987) Cohort mortality study of chemical workers with potential exposure to the higher chlorinated dioxins. J. Occup. Med., 29, 422–429.

Ott, M.G., Messerer, P. & Zober, A. (1993) Assessment of past occupational exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin using blood lipid analyses. Int. Arch. Occup. Environ. Health., 65, 1–8.

Ott, M.G., Zober, A. & Germann, C. (1994) Laboratory results for selected target organs in 138 individuals occupationally exposed to TCDD. Chemosphere, 29, 2423–2437.

van Overmeire, I., Goeyens, L., Beernaert, H., Srebrnik, S., De Poorter, G., Bayens, W., Clark, G., Chu, M., Chu, A., Morris, R. & Brown, D. (2000) A comparative study of GC/HRMS and CaluxTM TEQ determinations in food samples by the Belgian Federal Ministries of Public Health and Agriculture. Organohalogen Compd, 45, 196–199.

Päpke, O. (1998) PCDD/PCDF: Human background data for Germany, a 10 year experience. Environ. Health Perspectives, 106, 723–731.

Päpke, O. (1999) Lessons learned on performing more than 15,000 dioxin analyses. Organohalogen Compd, 44, 5–8.

Päpke, O., Ball, M., Lis, Z.A. & Scheunert, K. (1989) PCDD/PCDF in whole blood samples of unexposed persons. Chemosphere, 19, 941–948.

Päpke, O., Ball, M., Lis, A. & Wuthe, J. (1996) PCDD/PCDFs in humans, follow up of background data of Germany, 1994. Chemosphere, 32, 575–582.

Patandin, S., Koopman-Esseboom, C., de Ridder, M.A.J., Weisglas-Kuperus, N. & Sauer, P.J.J. (1998) Effects of environmental exposure to polychlorinated biphenyls and dioxins on birth size and growth in Dutch children. Pediatr. Res., 44, 538–545.

Patandin, S., Lanting, C.I., Mulder, P.G.H., Boersma, E.R., Sauer, P.J.J. & Weisglas-Kuperus, N. (1999) Effects of environmental exposure to polychlorinated biphenyls and dioxins on cognitive abilities in Dutch children at 42 months of age. J. Pediatr., 134, 33–41.

Patterson, D.G., Needham, L.L., Pirkle, J.L., Roberts, D.W., Bagby, J., Garrett, W.A., Andrews, J.S., Falk, H., Bernert, J.T., Sampson, E.J. & Houk, V.N. (1988) Correlation between serum and adipose tissue levels of 2,3,7,8-tetrachlorodibenzo-p-dioxin in 50 persons from Missouri. Arch. Environ. Contam. Toxicol., 17, 139–143.

Patterson, D.G., Fürst, P., Henderson, L.O., Isaacs, S.G., Alexander, L.R., Turner, W.E., Needham, L.L. & Hannon, H. (1989) Partitioning of in vivo bound PCDDs/PCDFs among various compartments in whole blood. Chemosphere, 19, 135–142.

Paumgartten F.J.R., Cruz, C.M., Chahoud, I., Palavinskas, R. & Mathar, W. (2000) PCDDs, PCDFs, PCBs and other organochlorine compounds in human milk from Rio de Janeiro, Brazil. Environ. Res., 83, 293–297.

Pauwels, A., Cenjin, P., Covaci, A., Weyler, J., Schepens, P.J.C. & Brouwer, A. (1999) Analysis of PCB congeners (by GC–ECD) and dioxin-like toxic equivalence (by CALUX assay) in females with endometriosis and other fertility problems. Organohalogen Compd, 44, 407–410.

Pauwels, A., Cenijn, P.H., Schepens, P.J. & Brouwer, A. (2000) Comparison of chemical- activated luciferase gene expression bioassay and gas chromatography for PCB determination in human serum and follicular fluid. Environ. Health Perspectives, 108, 553–557.

Pesatori, A.C., Consonni, D., Tironi, A., Zocchetti, C., Fini, A. & Bertazzi, P.A. (1993) Cancer in a young population in a dioxin-contaminated area. Int. J. Epidemiol., 22, 1010–1013.

Pesatori, A.C., Zocchetti, C., Guercilena, S., Consonni, D., Turrini, D. & Bertazzi, P.A. (1998) Dioxin exposure and non-malignant health effects: A mortality study. Occup. Environ. Med., 55, 126–131.

Peterson, R.E., Theobald, H.M. & Kimmel, G.L.. (1993) Developmental and reproductive toxicity of dioxins and related compounds: Cross-species comparisons. Crit. Rev. Toxicol., 23, 283–335.

Petroske, E., Zaylskie, R.G. & Feil, V.J. (1997) The effect of cooking on dioxin and furan concentrations in beef. Organohalogen Compd, 33, 436–439.

Petroske. E., Zaylskie, R.G. & Feil, V.J. (1998) Reduction in polychlorinated dibenzodioxin and dibenzofuran residues in hamburger meat during cooking. J. Agric. Food Chem., 46, 3280–3284.

Pinsky, P. & Lorber, M.N. (1998) A model to evaluate past exposure to 2,3,7,8-TCDD. J. Exposure Anal. Environ. Epidemiol., 8, 187–206.

van der Plas, S.A., Lutkeschipholt, I., Spenkelink, B. & Brouwer, A. (2001) Effects of subchronic exposure to complex mixtures of dioxin-like and non-dioxin-like polyhalogenated aromatic compounds on thyroid hormone and vitamin A levels in female Sprague-Dawley rats. Toxicol. Sci., 59, 92–100.

Pluess, N., Poiger, H. & Schlatter, C. (1987) The metabolism of some entachlorodibenzofurans in the rat. Xenobiotica, 17, 209–216.

Pluim, H.J., Wever, J., Koppe, J.G., van den Slikke, J.W. & Olie, K. (1993) Intake and faecal excretion of chlorinated dioxins and dibenzofurans in breast-fed infants at different ages. Chemosphere, 26,1947–1952.

Pocchiari, F., Silvano, V., Zampieri, A. & Zampieri, A. (1979) Human health effects from accidental release of tetrachlorodibenzo-p-dioxin (TCDD) at Seveso, Italy. Ann. N.Y. Acad. Sci., 77, 311–320.

Pohjanvirta, R., Vartianen, T., Uusi-Rauva, A., Mönkkönen, J. & Tuomisto, J. (1990) Tissue distribution, metabolism, and excretion of 14C-TCDD in a TCDD-susceptible and a TCDD resistant rat strain. Pharmacol. Toxicol., 66, 93–100.

Pohjanvirta, R., Wong, J., Li, W., Harper, P.A., Tuomisto, J. & Okey, A.B. (1998) Point mutation in intron sequence causes altered carboxyl-terminal structure in the aryl hydrocarbon receptor of the most 2,3,7,8-tetrachlorodibenzo-p-dioxin-resistant rat strain. Mol. Pharmacol., 54, 86–93.

Poiger, H. & Schlatter, C. (1986) Pharmacokinetics of 2,3,7,8-TCDD in man. Chemosphere, 15, 1489–1494.

Poiger, H., Pluess, N. & Buser, H.R. (1989a) The metabolism of selected PCDFs in the rat. Chemosphere, 18, 259–264.

Poiger, H., Pluess, N. & Schlatter, C. (1989b) Subchronic toxicity of some chlorinated dibenzofurans in rats. Chemosphere, 18, 265–275.

Poland, A., Teitelbaum, P. & Glover, E. (1989a) [125I]2-Iodo-3,7,8-trichlorodibenzo-p-dioxin-binding species in mouse liver induced by agonists for the Ah-receptor: Characterisation and identification. Mol. Pharmacol., 36, 113–120.

Poland, A., Teitelbaum, P., Glover, E. & Kende, A. (1989b) Stimulation of in vivo hepatic uptake and in vitro hepatic binding of [125I]2-iodo-3,7,8-trichlorodibenzo-p-dioxin by the administration of agonists for the Ah-receptor. Mol. Pharmacol, 36, 121–127.

Pollitt, F. (1999) Polychlorinated dibenzodioxins and polychlorinated dibenzofurans. Reg. Toxicol. Pharmacol., 30, 63–68.

Putzrath, R.M. (1997) Estimating relative potency for receptor-mediated toxicity: Reevaluating the toxicity equivalence factor (TEF) model. Regul. Toxicol. Pharmacol., 25, 68–78.

Rappe, C., Bergek, S., Fiedler, H. & Cooper, K.R. (1998a) PCDD and PCDF contamination in catfish feed from Arkansas, USA. Chemosphere, 36, 2705–2720.

Reggiani, G. (1980) Acute human exposure to TCDD in Seveso, Italy. J. Toxicol. Environ. Health, 6, 27–43.

Renwick, A.G. & Lazarus, N.R. (1998) Human variability and noncancer risk assessment. An analysis of the default uncertainty factor. Regul. Toxicol. Pharmacol., 27, 3–20.

Rhile, M.J., Nagarkatti, M. & Nagarkatti, P.S. (1996) Role of Fas apoptosis and MHC genes in 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)-induced immunotoxicity of T cells. Toxicology, 110, 153–167.

Ribas, N., Sala, M., Kogevinas, M. & Sunyer, J. (2001) Polychlorinated biphenyls and neurological development in children: A systematic review. J. Epidemiol. Commun. Health (in press).

Rier, S.E., Martin, D.C., Bowman, R.E., Dmowski, W.P. & Becker, J.L. (1993) Endometriosis in rhesus monkeys (Macaca mulatta) following chronic exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin. Fundam. Appl. Toxicol., 21, 433–441.

Rier, S.E., Turner, W.E., Martin, D.C., Morris, R., Lucier, G.W. & Clark, G.C. (2001a) Serum levels of TCDD and dioxin-like chemicals in rhesus monkeys chronically exposed to dioxin: Corrrelation of increased serum PCB levels with endometriosis. Toxicol. Sci., 59, 147–159.

Rier, S.E., Coe, C.L., Lemieux, A.M., Martin, D.C., Morris, R., Lucier, G.W. & Clark, G.C. (2001b) Increased tumor necrosis factor-a production by peripheral blood leukocytes from TCDD-exposed rhesus monkeys. Toxicol. Sci., 60, 327–337.

Roegner, R.H., Grubbs, W.D., Lustik, M.B., Brockman, A.S., Henderson, S.C., Williams, D.E., Wolfe, W.H., Michalek, J.E. & Miner, J.C. (1991) Air Force Health Study: An epidemiologic investigation of health effects in Air Force personnel following exposure to herbicides. Serum dioxin analysis of 1987 examination results (NTIS No. AD A-237-516 through AD A-237-524).

Rogan, W.J., Gladen, B.C., Hung, K.L., Koong, S.L., Shih, L.Y., Taylor, J.S., Wu, Y.C.,Yang, D., Ragan, N.B. & Hsu, C.C. (1988) Congenital poisoning by polychlorinated biphenyls and their contaminants in Taiwan. Science, 241, 334–336.

Rogan, W.J., Gladen, B.C., Guo, Y.-L.L. & Hsu, C.-C. (1999) Sex ratio after exposure to dioxin-like chemicals in Taiwan. Lancet, 353, 206–207.

Rogers, A.M., Andersen, M.E. & Back, K.C. (1982) Mutagenicity of 2,3,7,8-tetrachlorodibenzo-p-dioxin and perfluoro-n-decanoic acid in L5178Y mouse lymphoma cells. Mutat. Res., 105, 445–449.

Rohde, S., Moser, S., Päpke, O. & McLachlan, M.S. (1999) Clearance of PCDD/Fs via the gastrointestinal tract in occupationally exposed persons. Chemosphere, 38, 3397–3410.

Roman, B.L. & Peterson, R.E. (1998) Developmental male reproductive toxicology of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) and PCBs. Reprod. Dev. Toxicol., 24, 593–624.

Rose, J.Q., Ramsey, J.C., Wentzler, T.H., Hummel, R.A. & Gehring, P.J. (1976) The fate of 2,3,7,8-tetrachlorodibenzo-p-dioxin following single and repeated oral doses to the rat. Toxicol. Appl. Pharmacol., 36, 209–226.

Rozman, K.K. (1999) Delayed acute toxicity of 1,2,3,4,6,7,8-heptachlorodibenzo-p-dioxin (HpCDD), after oral administration, obeys Haber’s rule of inhalation toxicology. Toxicol. Sci., 49, 102–109.

Rozman, K.K. (2000) The role of time in toxicology or Haber’s c x t product. Toxicology, 149, 35–42.

Ryan, J.J. & Masuda, Y. (1989) Half-lives for elimination of polychlorinated dibenzofurans (PCDFs) and PCBs in humans from the Yusho and Yucheng rice oil poisonings. In: Proceedings of 9th International Symposium on Chlorinated Dioxins and Related Compounds, Dioxin ’89, 17–22 September 1989, Toronto, Ontario.

Ryan, J.J. & Masuda. Y (1994) Polychlorinated naphthalenes (PCNs) in the rice oil poisonings. Organohalogen Compd, 21, 251–254.

Ryan, J.J., Lizotte, R. & Lau, B.P.Y. (1985) Chlorinated dibenzo-p-dioxins and chlorinated dibenzofurans in Canadian human adipose tissue. Chemosphere, 14, 697–706.

Ryan, J.J., Gasiewicz, T.A. & Brown, J.F. (1990) Human body burden of polychlorinated dibenzofurans associated with toxicity based on the Yusho and Yucheng incidents. Fundam. Appl. Toxicol., 15, 722–731.

Ryan, J.J., Levesque, D., Panopio, L.G., Sun, W.F., Masuda, Y. & Kuroki, H. (1993a) Elimination of polychlorinated dibenzofurans (PCDFs) and polychlorinated biphenyls (PCBs) from human blood in the Yusho and Yu-cheng rice oil poisonings. Arch. Environ. Contam. Toxicol., 24, 504–512.

Ryan, J.J., Lizotte, R., Panopio, L.G., Shewchuk, C., Lewis, D.A. & Sun, W.F. (1993) Polychlorinated dibenzo-p-dioxins (PCDDs) and polychlorinated dibenzofurans (PCDFs) in human milk samples collected across Canada in 1986–87. Food Addit. Contam., 10, 419–428.

Ryan, J.J., Hsu, C.-C., Boyle, M.J. & Guo, Y.-L.L. (1994) Blood serum levels of PCDFs and PCBs in Yu-cheng children post-natally exposed to toxic rice oil. Chemosphere, 29, 1263–1278.

Rylander, L., Stromberg, U. & Hagmar, L. (1995) Decreased birthweight among infants born to women with a high dietary intake of fish contaminated with persistent organochlorine compounds. Scand. J. Work Environ. Health, 21, 368–375.

Santastefano, M.J., Johnson, K.L., Whisnant, N.A., Richardson, V.M., DeVito, M.J., Diliberto, J.J. & Birnbaum, L.S. (1996) Subcellular localization of TCDD differs between the liver, lungs and kidneys after acute and subchronic exposure: Species/dose comparisons and possible mechanism. Fundam. Appl. Toxicol., 34, 265–275.

Saracci, R., Kogevinas, M., Bertazzi, P.A., Bueno de Mesquita, B.H., Coggon, D., Green, L.M., Kauppinen, T., L’Abbé, K.A., Littorin, M., Lynge, E., Mathews, J.D., Neuberger, M., Osman, J., Pearce, N. & Winkelmann, R. (1991) Cancer mortality in workers exposed to chlorophenoxy herbicides and chlorophenols. Lancet, 338, 1027–1032.

Schantz, S.L. & Bowman, R.E. (1989) Learning in monkeys exposed perinatally to 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD). Neurotoxicol. Teratol., 11, 13–19.

Schantz, S.L, Ferguson, S.A. & Bowman, R.E. (1992) Effects of 2,3,7,8-tetrachlorodi-benzo-p-dioxin on behavior of monkeys in peer groups. Neurotoxicol. Teratol., 14, 433–446.

Schecter, A., Ryan, J.J., Constable, J.D., Baughman, R., Bangert, J., Furst, P., Wilmers, K. & Oates, R.P. (1990) Partitioning of 2,3,7,8-chlorinated dibenzo-p-dioxins and dibenzofurans between adipose tissue and plasma lipid of 20 Massachusetts Vietnam veterans. Chemosphere, 20, 951–958.

Schecter, A., Päpke, O., Ball, M. & Ryan, J. (1991) Partitioning of dioxins and dibenzofurans: Whole blood, blood plasma and adipose tissue. Chemosphere, 23, 1913–1919.

Schecter, A., Päpke, O., Ball, M., Lis, A. & Brandt-Rauf, P. (1995) Dioxin concentrations in the blood of workers at municipal waste incinerators. Occup. Environ. Med., 52, 385–387.

Schecter, A.J., Päpke, O., Dellarco, M. & Olson, J.R. (1996) A comparison of dioxins and dibenzofurans in cooked and uncooked food. Organohalogen Compd, 28, 166–170.

Schecter, A., Päpke, O & Dellarco, M. (1997) Dioxin, dibenzofuran, and PCB congeners in cooked and uncooked foods. Organohalogen Compd, 33, 462–466.

Schecter, A., Kassis, I. & Päpke, O. (1998a) Partitioning of dioxins, dibenzofurans, and coplanar PCBs in blood, milk, adipose tissue, placenta and cord blood from five American women. Chemosphere, 37, 1817–1823.

Schecter, A., Dellarco, M., Päpke, O. & Olson, J. (1998b) A comparison of dioxins, dibenzofurans and coplanar PCBs in uncooked and broiled ground beef, catfish and bacon. Chemosphere, 37, 1723–1730.

Schlatter, C. (1991) Data on kinetics of PCDDs and PCDFs as a prerequisite for human risk assessment. In: Biological Basis for Risk Assessment of Dioxin and Related Compounds, Cold Spring Harbor, NY: CSH Press, pp. 215–226.

Schrenk, D., Lipp, H.-P., Wiesmüller, T., Hagenmaier, H. & Bock, K.W. (1991) Assessment of biological activities of mixtures of polychlorinated dibenzo-p-dioxins: Comparison between defined mixtures and their constituents. Arch. Toxicol., 65, 114–118.

Schrenk, D., Buchmann, A., Dietz, K., Lipp, H.P., Brunner, H., Sirma, H., Münzel, P., Hagenmaier, H., Gebhardt, R. & Bock, K.W. (1994) Promotion of preneoplastic foci in rat liver with 2,3,7,8-tetrachlorodibenzo-p-dioxin, 1,2,3,4,6,7,8-heptachlorodibenzo-p-dioxin and a defined mixture of 49 polychlorinated dibenzo-p-dioxins. Carcinogenesis, 15, 509–515.

Schrenk, D., Stüven, T., Gohl, Viebahn, R. & Bock, K.W. (1995) Induction of CYP1A and glutathione S-transferase activities by 2,3,7,8-tetrachlorodibenzo-p-dioxin in human hepatocyte cultures. Carcinogenesis, 16, 943–946.

Schulte, E. & Malisch, R. (1983) Berechnung der wahren PCB-Gehalte in Umweltproben. I. Ermittlung der Zusammensetzung zweier technischer PCB-Gemische. Fresenius Z. Anal. Chem., 314, 545–551.

Schuur, A.G., Boekhorst, F.M., Brouwer, A. & Visser, T.J. (1997) Extrathyroidal effects of 2,3,7,8-tetrachlorodibenzo-p-dioxin on thyroid hormone turnover in male Sprague-Dawley rats. Endocrinology, 138, 3727–3734.

Schwetz, B.A., Norris, J.M., Sparschu, G.L., Rowe, U.K., Gehring, P.J., Emerson, J.L. & Gerbig, C.G. (1973) Toxicology of chlorinated dibenzo-p-dioxins. Environ. Health Perspectives, 5, 87–99.

Schwind, K.H. & Hecht, H. (2000) Mitteilungsblatt No. 149, September 2000, Bundesministerium für Ernährung, Landwirtschaft und Forsten, Bundesanstalt für Fleischforschung, 39, 803–806.

Seo, B.-W., Li, M.-H., Hansen, L.G., Moore, R.W., Peterson, R.E. & Schantz, S.L. (1995) Effects of gestational and lactational exposure to polychlorinated biphenyl (PCB) congeners or 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) on thyroid hormone concentrations in weanling rats. Toxicol. Lett., 78, 253–262.

Sewall, C.H., Flagler, N., Van den Heuvel, J.P., Clark, G.C., Tritscher, A.M., Maronpot, R.M. & Lucier, G.W. (1995) Alterations in thyroid-function in female Sprague-Dawley rats following chronic treatment with 2,3,7,8-tetrachlorodibenzo-p-dioxin. Toxicol. Appl. Pharmacol., 132, 237–244.

Slob, W. (1993) Modelling long-term exposure of the whole population to chemicals in food. Risk Anal., 13, 525–530.

Smialowicz, R.J., Riddle, M.M., Williams, W.C. & Diliberto, J.J. (1994) Effects of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) on humoral immunity and lymphocyte subpopulations: Differences between mice and rats. Toxicol. Appl. Pharmacol., 124, 248–256.

Smialowicz, R.J., Williams, W.C. & Riddle, M.M. (1996) Comparison of the T cell-independent antibody response of mice and rats exposed to 2,3,7,8-tetrachlorodibenzo-p dioxin. Fundam. Appl. Toxicol., 32, 293–297.

Smith, A.H., Patterson, D.G., Jr, Warner, M.L., Mackenzie, R. & Needham, L.L. (1992) Serum 2,3,7,8-tetrachloro-dibenzo-o-dioxin levels of New Zealand pesticide applicators and their implication for cancer hypotheses. J. Natl Cancer Inst., 84, 104–108.

Smits-van Prooije, A.E., Lammers, J.H.C.M., Waalkens-Berendsen, D.H., Kulig, B.M. & Snoeij, N.J. (1993) Effects of the PCB 3,4,5,3,4,5,hexachlorobiphenyl on the reproduction capacity of Wistar rats. Chemosphere, 27, 395–400.

Stahl, B.U., Kettrup. A. & Rozman, K. (1992) Comparative toxicity of four chlorinated dibenzo-p-dioxins (CDDs) and their mixture. Part I: Acute toxicity and toxic equivalency factors (TEFs). Arch. Toxicol., 66, 471–477.

Steenland, K., Piacitelli, L., Deddens, J., Fingerhut, M. & Chang, L.I. (1999) Cancer, heart disease, and diabetes in workers exposed to 2,3,7,8-tetrachlorodibenzo-p-dioxin. J. Natl Cancer Inst., 91, 779–786.

Steenland, K., Deddens, J. & Piacitelli, L. (2001) Risk assessment for 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) based on an epidemiologic study. Am. J. Epidemiol., 154, 451–458.

Stonard, M.D. & Greig, J.B. (1976) Different patterns of hepatic microsomal enzyme activity produced by administration of pure hexachlorobiphenyl isomers and hexachlorobenzene. Chem.–Biol. Interactions, 15, 365–379.

Suskind, R.R. & Hertzberg, V.S. (1984) Human health effects of 2,4,5-T and its toxic contaminants. J. Am. Med. Assoc., 251, 2372–2380.

Sweeney, M.H. & Mocarelli, P. (2000) Human health effects after exposure to 2,3,7,8-TCDD. Food Addit. Contam., 17, 303–316.

Tai, H.-L., McReynolds, J.H., Goldstein, J.A., Eugster, H.P., Sengtag, C., Alworth, W.L. & Olson, J.A. (1993) Cytochrome P450 1A1 mediates the metabolism of 2,3,7,8-tetrachlorodibenzofuran in the rat and human. Toxicol. Appl. Pharmacol., 123, 34–42.

Tanabe, S., Kannan, N., Wakimoto, T., Tatsukawa, R., Okamoto, T. & Masuda, Y. (1989) Isomer-specific determination and toxic evaluation of potentially hazardous coplanar PCBs, dibenzofurans and dioxins in the tissues of "Yusho" PCB poisoning victim and in the causal oil. Toxicol. Environ. Chem., 24, 215–231.

Tanaka, N., Nettesheim, P., Gray, T., Nelson, K. & Barrett, J.C. (1989) 2,3,7,8-Tetrachlorodi-benzo-p-dioxin enhancement of N-methyl-N’-nitro-N-nitrosoguanidine induced transformation of rat tracheal epithelial cells in culture. Cancer Res., 49, 2703–2708.

Tenchini, M.L., Crimaudo, C., Pacchetti, G., Mottura, A., Agosti, S. & DeCarli, L. (1983) A comparative cytogenetic study on cases of induced abortions in TCDD exposed and non-exposed women. Environ. Mutag., 5, 73–85.

Theobald, H.M. & Peterson, R.E. (1997) In utero and lactational exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin: Effects on development of the male and female reproductive system of the mouse. Toxicol. Appl. Pharmacol., 145, 124–135.

Thiel, R., Koch, E., Ulbrich, B. & Chahoud, I. (1994) Peri- and postnatal exposure to 2,3,7,8- tetrachlorodibenzo-p-dioxin: Effects on physiological development, reflexes, locomotor activity and learning behaviour in Wistar rats. Arch. Toxicol., 69, 79–86.

Thiess, A.M., Frentzel-Beyme, R. & Link, R. (1982) Mortality study of persons exposed to dioxin in a trichlorophenol-process accident that occurred in the BASF AG on November 17, 1953. Am. J. Ind. Med., 3, 179–189.

Thigpen, J.E., Faith, R.E., McConnell, E.E. & Moore, J.A. (1975) Increased susceptibility to bacterial infection as a sequela of exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin. Infect. Immun., 12, 1319–1324.

Thoma, H., Mücke, W. & Kauert, G. (1990) Comparison of the polychlorinated dibenzo-p-dioxin and dibenzo-p-furan in human tissue and human liver. Chemosphere, 20, 433–442.

Thomas, P.T. & Hinsdill, R.D. (1979) The effect of perinatal exposure to tetrachlorodibenzo-p-dioxin on the immune response of young mice. Drug Chem. Toxicol., 2, 77–98.

Thornton, A.S., Oda, Y., Stuart, G.R., Glickman, B.W. & De Boer, J.G. (2001) Mutagenicity of TCDD in Big Bleu® transgenic rats. Mutat. Res. (in press).

Thorpe et al. (1999) Organohalogen Compd, 44, 237–240.

Tóth, K., Somfai-Relle, S., Sugár, J. & Bence, J. (1979) Carcinogenicity testing of herbicide 2,4,5-trichlorophenoxyethanol containing dioxin and of pure dioxin in Swiss mice. Nature, 278, 548–549.

Townsend, J.C., Bodner, K.M., Van Peenen, P.F., Olson, R.D. & Cook, R.R. (1982) Survey of reproductive events of wives of employees exposed to chlorinated dioxins. Am. J. Epidemiol., 115, 695–713.

Traag, W.A., Mengelers, M.J.B., Kan, C.A. & Malisch, R. (1999) Studies on the uptake and carry over of polychlorinated dibenzodioxins and dibenzofurans from contaminated citrus pulp pelllets to cows milk. Organohalogen Compd, 42, 201–204.

Tritscher, A.M., Goldstein, J.A., Portier, C.J., McCoy, Z. & Clark, G.C. (1992) Dose–response relationships for chronic exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin in a tumor promotion model: Quantification and immunolocalization of CYP1A1 and CYP1A2 in the liver. Cancer Res., 52, 3436–3442.

Tucker, A.N., Vore, S.J. & Luster, M.I. (1986) Suppression of B cell differentiation by 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) on the thymus. Toxicology, 15, 135–144.

Tuey, D.B. & Matthews, H.B. (1980) Use of a physiological compartmental model for the rat to describe the pharmacokinetics of several chlorinated biphenyls in the mouse. Drug Metab. Disposition, 8, 397–403.

Turteltaub, K.W., Felton, J.S., Gledhill, B.L., Vogel, J.S., Southon, J.R., Caffee, M.W., Finkel, R.C., Nelson, D.E., Proctor, I.D. & Davis, J.C. (1990) Accelerator mass spectroscopy in biomedical dosimetry: Relationship between low-level exposure and covalent binding of heterocyclic amine carcinogens to DNA. Proc. Natl Acad. Sci. USA, 87, 5288–5292.

UNEP (1999) Dioxin and Furan Inventories: National and Regional Emissions of PCDD/PCDF, Geneva

UNEP (2001) Standardized Toolkit for Identification and Quantification of Dioxin and Furan Releases, Geneva

Van Miller, J.P., Lalich, J.J. & Allen, J.R. (1977) Increased incidence of neoplasms in rats exposed to low levels of 2,3,7,8-tetrachlorodibenzo-p-dioxin. Chemosphere, 10, 625–632.

Vecchi, A., Sironi, M., Canegrati, M.A., Recchia, M. & Garattini, S. (1983) Immunosuppressive effects of 2,3,7,8-tetrachlorodibenzo-p-dioxin in strains of mice with different susceptibility to induction of aryl hydrocarbon hydroxylase. Toxicol. Appl. Pharmacol., 68, 434–441.

Vena, J., Boffetta, P., Becher, H., Benn, T., Bueno de Mesquita, H.B., Coggon, D., Colin, D., Flesch-Janys, D., Green, L., Kauppinen, T., Littorin, M., Lynge, E., Mathews, J.D., Neuberger, M., Pearce, N., Pesatori, A.C., Saracci, R., Steenland, K. & Kogevinas, M. (1998) Exposure to dioxin and non-neoplastic mortality in the expanded IARC international cohort study of phenoxy herbicide and chlorophenol production workers and sprayers. Environ. Health Perspectives, 106 (Suppl. 2), 645–653.

Viluksela, M., Duong, T.V., Stahl, B.U., Li, X., Tuomisto, J. & Rozman, K.K. (1996) Toxicokinetics of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in two substrains of male Long-Evans rats after intraveneus injection. Fundam. Appl. Toxicol., 31, 184–191.

Viluksela, M., Stahl, B.U., Birnbaum, L.S., Schramm, K.W., Kettrup, A. & Rozman, K.K. (1997) Subchronic/chronic toxicity of 1,2,3,4,6,7,8-heptachlorodibenzo-p-dioxin (HpCDD) in rats. Toxicol. Appl. Pharmacol., 146, 207–226.

Viluksela, M., Stahl, B.U., Birnbaum, L.S., Schramm, K.-W., Kettrup, A. & Rozman, K.K. (1998a) Subchronic/chronic toxicity of a mixture of four chlorinated dibenzo-p-dioxins in rats. I. Design, general observations, hematology, and liver concentrations. Toxicol. Appl. Pharmacol., 151, 57–69.

Viluksela, M., Stahl, B.U., Birnbaum, L.S. & Rozman, K.K. (1998b) Subchronic/chronic toxicity of a mixture of four chlorinated dibenzo-p-dioxins in rats. II. Biochemical effects. Toxicol. Appl. Pharmacol., 151, 70–78.

Voedingscentrum (1998) Zo eet Nederland 1998, The Hague (in Dutch).

Vos, J.G. & Moore, J.A. (1974) Suppression of cellular immunity in rats and mice by maternal treatment with 2,3,7,8-tetrachlorodibenzo-p-dioxin. Int. Arch. Allergy, 47, 777–794.

Vos, J.G., Moore, J.A. & Zinkl, J.G. (1973) Effect of 2.3,7,8-tetrachlorodibenzo-p-dioxin on the immune system of laboratory animals. Environ. Health Perspectives, 5, 149–162.

Vos, J.G., Moore, J.A. & Zinkl, J.G. (1974) Toxicity of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in C57B1/6 mice. Toxicol. Appl. Pharmacol., 29, 229–241.

Wacker, R., Poiger, H. & Schlatter, C. (1986) Pharmacokinetics and metabolism of 1,2,3,7,8-pentachlorodibenzo-p-dioxin in the rat. Chemosphere, 15, 1473.

Wahba, Z.Z., Lawson, T.A. & Stohs, S.J. (1988) Induction of hepatic DNA single strand breaks in rats by 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD). Cancer Lett., 39, 281–286.

Wahba, Z.Z., Lawson, T.W., Murray, W.J. & Stohs, S.J. (1989) Factors influencing the induction of DNA single strand breaks in rats by 2,3,7,8-tetrachlorodibenzo-p-dioxin. Toxicology, 58, 57–69.

Wang, X., Santastefano, M.J., Evans, M.A., Richardson, V.M., Diliberto, J.J. & Birnbaum, L.S. (1997a) Determination of parameters responsible of pharmacokinetic behavior of TCDD in female Sprague-Dawley rats. Toxicol. Appl. Pharmacol., 147, 151–168.

Wang, X., Santastefano, M.J., DeVito, M.J. & Birnbaum, L.S. (1997b) Extrapolation of a previous PBPK model for TCDD across routes of exposure, genders, and from rats to mice. Organohalogen Compd, 34, 38–42.

Watson, M.A., Devereux, T.R., Malarkey, D.E., Andersen, M.W. & Maronpot, R.R. (1995) H-ras oncogene mutation spectra in B6C3F1 and C57BL/6 mouse liver tumors provide evidence for TCDD promotion of spontaneous and vinyl carbamate initiated liver cells. Carcinogenesis, 16, 1705–1710.

Weber, L.W.D., Zesch, A. & Rozman, K. (1991) Penetration, distribution and kinetics of 2,3,7,8-tetrachloro-p-dioxin in human skin in vitro. Arch. Toxicol., 65, 421–428.

Weber, L.W.D., Ernst, S.W., Stahl, B.U. & Rozman, K. (1993) Tissue distribution and toxicokinetics of 2,3,7,8-tetrachlorodibenzo-p-dioxin in rats after intravenous injection. Fundam. Appl. Toxicol., 21, 523–534.

Weisglas-Kuperus, N., Patandin, S., Berbers, G.A.M., Sas, T.C.J., Mulder, P.G.H., Sauer, P.J.J. & Hooijkaas, H. (2000) Immunologic effects of background exposure to polychlorinated biphenyls and dioxins in Dutch preschool children. Environ. Health Perspectives, 108, 1203–1207.

Wendling, J.M., Orth, R.G. & Poiger, H. (1990) Determination of [3H]-2,3,7,8-tetrachlorodibenzo-p-dioxin in human feces to ascertain its relative metabolism in man. Anal. Chem., 62, 796–800.

Whitlock, J.P., Chichester, C.H., Bedgood, R.M., Okino, S.T., Ko, H.P., Ma, Q., Dong, L., Li, H. & Clarke-Katzenberg, R. (1997) Induction of drug-metabolizing enzymes by dioxin. Drug Metab. Rev., 29, 1107–1127.

WHO (1987) Principles for the Safety Assessment of Food Additives and Contaminants in Food (Environmental Health Criteria 70), Geneva.

WHO (1994) Assessing Human Health Risks of Chemicals: Derivation of Guidance Values for Health-based Exposure Limits (Environmental Health Criteria 170), Geneva.

WHO (1995) Quality Assessment of PCBs, PCDDs and PCDF Analyses: Third Round of WHO-coordinated Study, Copenhagen, WHO Regional Office for Europe.

WHO (1996) Levels of PCBs, PCDDs and PCDFs in Human Milk (Environmental Health in Europe No. 3), Bilthoven: WHO European Centre for Environment and Health.

WHO (1998) GEMS/Food Regional Diets (WHO/FSF/FOS/98.3), GEMS/Food Programme, Geneva.

WHO (2000) Methodology for Exposure Assessment of Contaminants and Toxins in Food, Report of a Joint FAO/WHO Workshop, Geneva, Switzerland, 7–8 June 2000 (WHO/SDE/PHE/FOS/00.5), Geneva.

Wilson, J.D. (1987) A dose–response curve for Yusho syndrome. Regul. Toxicol. Pharmacol., 7, 364–369.

Wittsiepe, J., Schrey, P., Ewers, U., Selenka, F. & Wilhelm, M. (2000) Decrease of PCDD/F levels in human blood—Trend analysis for the German population, 1991–1996. Environ. Res. A, 83, 46–53.

Wolf, C.J., Ostby, J.S. & Gray, L.E., Jr (1999) Gestational exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) severely alters reproductive function in female hamster offspring. Toxicol. Sci., 51, 259–265.

Wolfe, W.H., Michalek, J.E., Miner, J.C., Pirkle, J.L., Caudill, S.P., Patterson, D.G. & Needham, L.L. (1994) Determinants of TCDD half-life in veterans of Operation Ranch Hand. J. Toxicol. Environ. Health, 41, 481–488.

Wolfe, W.H., Michalek, J.E., Miner, J.C., Rahe, A.J., Moore, C.A., Needham, L.L. & Patterson, D.G., Jr (1995) Paternal serum dioxin and reproductive outcomes among veterans of Operation Ranch Hand. Epidemiology, 6, 17–22.

Wroblewsky, V.J. & Olson, J.R. (1985) Hepatic metabolism of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in the rat and guinea pig. Toxicol. Appl. Pharmacol., 81, 231–240.

Yamashita, F. & Hayashi, M. (1985) Fetal PCB syndrome: Clinical features, intrauterine growth retardation and possible alteration in calcium metabolism. Environ. Health Perspectives, 59, 41–55.

Yang, J.H., Thraves, P., Dritschilo, A. & Rhim, J.S. (1992) Neoplastic transformation of immortalized human keratinocytes by 2,3,7,8-tetrachlorodibenzo-p-dioxin. Cancer Res., 52, 3478–3482.

Yang, J.Z., Agarwal, S.K. & Foster, W.G. (2000) Subchronic exposure to 2,3,7,8-tetrachlorodi-benzo-p-dioxin modulates the pathophysiology of endometriosis in the cynomolgus monkey. Toxicol. Sci., 56, 374–381.

Yao, Y., Takasuga, T., Masunaga, S. & Nakanishi, J. (2002) Detailed study on the levels of polychlorinated dibenzo-p-dioxins, polychlorinated dibenzofurans and polychlorinated biphenyls in Yusho rice oil. Chemosphere, 46, 1461–1469.

Yonemoto, J. (2000) The effects of dioxin on reproduction and development (review). Ind. Health, 38, 259–263.

Yoshimura, T. (1996) Growth of affected children. In: Kuratsune, M., Yoshimura, H., Hori, Y. & Okumura, M., eds, Yusho: A Human Disaster Caused by PCBs and Related Compounds, Fukuoka: Kyushu University Press, pp. 218–224.

Yoshimura, T. & Hayabuchi, H. (1985) Relationship between the amount of rice oil ingested by patients with Yusho and their subjective symptoms. Environ. Health Perspectives, 59, 47–51.

Yoshimura, H., Kuroki, J., Koga, N., Kuroki, H., Masuda, Y., Faukasaku, N. & Hasegawa, M. (1984) High accumulation of 2,3,4,7,8-pentachlorodibenzofuran to hepatic microsomes of rats. J. Pharmaco-Dyn., 7, 414–419.

Yoshimura, Y., Kamimura, H., Oguri, K., Honda, Y. & Nakano, M. (1986) Stimulating effects of activated charcoal beads on fecal excretion of 2,3,4,7,8-pentachlorodibenzofuran. Chemosphere, 15, 219–225.

Yu, M.-L., Guo, Y.-L.L., Rogan, W.J. & Hsu, C.-C. (1994) The 13-year mortality experience of the Yu-cheng cohort and their controls. Organohalogen Compd, 21, 489–493.

Yu, M.L., Guo, Y.L., Hsu, C.C. & Rogan, W.J. (2000) Menstruation and reproduction in women with polychlorinated biphenyl (PCB) poisoning: Long-term follow-up interviews of the women from the Taiwan Yu-cheng cohort. Int. J. Epidemiol., 29, 672–677.

Zack, J.A. & Gaffey, W.R. (1983) A mortality study of workers employed at the Monstanto Company plant in Nitro, West Virginia. In: Tucker, R., Young, A. & Grey, A., eds, Human and Environmental Risks of Chlorinated Dioxins and Related Compounds, New York, Plenum Press, pp . 575–591.

Zack, J.A. & Suskind, R.R. (1980) The mortality experience of workers exposed to tetrachlorodibenzodioxin in a trichlorophenol process accident. J. Occup. Med., 22, 11–14.

Zeiger, E., Anderson, B., Haworth, S., Lawlor, T. & Mortelmans, K. (1988) Salmonella mutagenicity tests: IV. Results from the testing of 300 chemicals. Environ. Mol. Mutag., 11 (Suppl. 12), 1–158.

Zeilmaker, M.J. & van Eijkeren, J. (1997) Modeling of Ah-receptor dependent P450 induction. I. Cellular model definition and its incorporation in a PBPK model of 2,3,7,8-TCDD (RIVM report 60138.001), Bilthoven: National Institute of Public Health and the Environment.

Zeilmaker, M.J. & van Eijkeren, J.C.H. (1998) The calculation of human toxicity thresholds of 2,3,7,8-TCDD: A physiologically based pharmacokinetic modeling approach (RIVM report 601503.010), Bilthoven: National Institute of Public Health and the Environment.

Zeilmaker, M.J., Fiolet, D. & Cuijpers, C.E.J. (1999) The contribution of hepatic protein binding on the body burden of dioxins in humans: A PBPK modeling analysis (RIVM report 529102.010), Bilthoven: National Institute of Public Health and the Environment (in Dutch with English abstract).

Zober, A., Messerer, P. & Huber, P. (1990) Thirty-four-year mortality follow-up of BASF employees exposed to 2,3,7,8-TCDD after the 1953 accident. Int. Arch. Occup. Environ. Health, 62, 139–157.

Zober, A., Ott, M.G., Fleig, I. & Heidemann, A. (1993) Cytogenetic studies in lymphocytes of workers exposed to 2,3,7,8-TCDD. Int. Arch. Occup. Environ. Health, 65, 157–161.

Zober, A., Ott, M.G. & Messerer, P. (1994) Morbidity follow up study of BASF employees exposed to 2,3,7, 8-tetrachlorodibenzo-p-dioxin (TCDD) after a 1953 chemical reactor incident. Occup. Environ. Med., 51, 479–486.

Zober, A., Messerer, P. & Ott, M.G. (1997) BASF studies: Epidemiological and clinical investigations on dioxin-exposed chemical workers. Teratog. Carcinog. Mutag., 17, 249–256.

Appendix 1. Procedure for calculating dietary intake of coplanar PCBs

Dietary intake

In order to calculate the distribution of the long-term mean intake of contaminants in a certain region, the long-term mean intake by an individual in the population of the region during a period T is represented by:

EQUATION 1

(1)

where J(line) T is the long-term mean personal intake of a contaminant (M/T); Ci(t) is the concentration of the contaminant in food group i (M/M); Ii(t)is the consumption of foods by group i (M/T); t is time (T); and N is the number of food groups considered.

If for each food group i we define: (i) a mean concentration for period T, C(line)i, (ii) a mean food consumption for period T, I(line)i, and (iii) temporal deviations from these means, ei(C(line)i,t) and ei(I(line)i,t), we can write:

EQUATION 2

(2)

and accordingly:

EQUATION 3

(3)

If the deviations from the mean food consumption and concentration are not correlated, for long averaging periods T, Eq. (3) reduces to:

EQUATION 4

(4)

and thus:

EQUATION 5

(5)

If we define the fraction of food group i, fi, that contributes to the mean food consumption, I(line)t then we can write:

EQUATION 6

(6)

The distribution of long-term mean intake in a certain region can now easily be found by perceiving long-term mean personal food consumption, mean contaminant concentrations in the food groups, and the fractional contribution of food groups to the total food consumption as random variables.

Basic assumptions about the random variables

For each concentration distribution and diet, we assume:

  • The vector of fractions of various food groups contributing to the total diet has a multinomial distribution. The uncertainty in the probability vector is assumed to approximate zero.
  • All other random variables are log-normally distributed.
  • In the variation of mean personal food consumption, we recognize between-person variation in food consumption distribution.
  • In the variation of concentrations, we recognize between-country variation within each region in concentration distribution.

Monte Carlo simulations

With the distributions specified above used as input data, the distributions of long-term mean intake can be estimated from Monte Carlo simulations.

Appendix 2. Non-compliance and mean in a truncated log-normal distribution

The probability density of a log-normal distribution is given by:

EQUATION 1

(1)

where sigmag is the geometric standard deviation (GSD) and mu is the median (or geometric mean). Consider a value xm (maximum limit) in this distribution. We want to know (i) the probability that this value is exceeded (non-compliance) and (ii) the mean of the distribution that appears when the original log-normal distribution is truncated at xm (see Figure A2).

FIGURE A2

Figure A2. Log-normal distribution and explanation of terms

Non-compliance

To answer the first question, the following integral must be solved:

EQUATION 2

(2)

where

EQUATION 3

(3)

Define

EQUATION 4

(4)

differencing yields:

EQUATION 5

(5)

Accordingly, (3) can be written as:

EQUATION 6

(6)

EQUATION 7

(7)

As the above integral corresponds to the definition of the error function, the solution of the definite integral is

EQUATION 8

(8)

EQUATION 9

(9)

Then, non-compliance is

EQUATION 10

(10)

Mean truncated distribution

The mean of the truncated distribution follows from:

EQUATION 11

(11)

For the solution of the integral in the denominator, see Eq. (9). Hence, the remaining integral to be solved is:

EQUATION 12

(12)

Use of the same substitution as in (3) yields

EQUATION 13

(13)

Define:

EQUATION 14

(14)

Thus:

EQUATION 15

(15)

for which the solution is (Abramovitz & Stegun, 1966):

EQUATION 16

(16)

Accordingly:

EQUATION 17

(17)

EQUATION 18

(18)

Reference

Abramowitz, M. & Stegun, I.A. (1966) Handbook of Mathematical Functions (Applied Mathematics Series 55). National Bureau of Standards, Washington DC,

ENDNOTES

1 This determination was based on a calculation made with a computer program graciously provided to the Committee by Dr Steenland.

2 By analogy with the provisional tolerable weekly intake (PTWI), the end-point used for safety evaluations by the Committee for food contaminants with cumulative properties. Its value represents the permissible human monthly exposure to these contaminants unavoidably associated with otherwise wholesome, nutritious foods.



    See Also:
       Toxicological Abbreviations