National Academies Press: OpenBook

Veterans and Agent Orange: Update 2008 (2009)

Chapter: 4 Information Related to Biologic Plausibility

« Previous: 3 Exposure to the Herbicides Used in Vietnam
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 65
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 66
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 67
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 68
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 69
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 70
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 71
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 72
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 73
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 74
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 75
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 76
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 77
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 78
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 79
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 80
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 81
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 82
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 83
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 84
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 85
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 86
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 87
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 88
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 89
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 90
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 91
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 92
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 93
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 94
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 95
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 96
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 97
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 98
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 99
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 100
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 101
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 102
Suggested Citation:"4 Information Related to Biologic Plausibility." Institute of Medicine. 2009. Veterans and Agent Orange: Update 2008. Washington, DC: The National Academies Press. doi: 10.17226/12662.
×
Page 103

Below is the uncorrected machine-read text of this chapter, intended to provide our own search engines and external engines with highly rich, chapter-representative searchable text of each book. Because it is UNCORRECTED material, please consider the following text as a useful but insufficient proxy for the authoritative book pages.

4 Information Related to Biologic Plausibility The committee reviewed all relevant experimental studies of 2,4-dichlo- rophenoxyacetic acid (2,4-D), 2,4,5-trichlorophenoxyacetic acid (2,4,5-T), pi- cloram, cacodylic acid, and 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) that have been published since Update 2006 (IOM, 2007) and has incorporated the findings, when it was appropriate, into this chapter or into the biologic-plausibil- ity sections of Chapters 6–9 when they are of consequence for particular health outcomes. For each substance, this chapter includes a review of toxicokinetic properties, a brief summary of the toxic outcomes investigated in animal experi- ments, and a discussion of underlying mechanisms of action as illuminated by in vitro studies. To achieve the goals of this chapter more effectively, the current committee has slightly modified the presentation of toxicologic information used by previous Veterans and Agent Orange (VAO) committees. The toxicology chapter of each earlier update presented information about each of the several hundred potentially relevant toxicologic articles published in the preceding 2 years. In contrast with the committee’s responsibility to evaluate each potentially relevant epidemiologic study of the chemicals of interest published in the preceding 2 years, its charge with respect to the toxicologic literature is to distill experimental toxicologic findings to judge whether it is biologically plausible to attribute adverse health outcomes reported in epidemiologic investigations to the chemicals. The current committee recognized that for most readers of the VAO series the implications of most toxicologic results reported are not immediately obvious. Therefore, start- ing with this update, the committee will focus on integrating and interpreting the toxicologic evidence rather than delineating the entire body of new experimental findings. 65

66 VETERANS AND AGENT ORANGE: UPDATE 2008 Establishment of biologic plausibility through laboratory studies strengthens the evidence of a cause–effect relationship between herbicide exposure and health effects reported in epidemiologic studies and thus supports the existence of the less stringent relationship of association, which is the target of this committee’s charge. Experimental studies of laboratory animals or cultured cells allow obser- vation of effects of herbicide exposure under highly controlled conditions that are difficult or impossible to control in epidemiologic studies. Such conditions include frequency and magnitude of exposure, exposure to other chemicals, pre- existing health conditions, and genetic differences between people, all of which can be controlled in a laboratory animal study. Once a chemical contacts the body, it begins to interact through the processes of absorption, distribution, metabolism, and excretion. Those four biologic pro- cesses characterize the disposition of a foreign substance that enters the organ- ism. Their combination determines the concentration of the compound in the body and how long each organ is exposed to it and thus influences its toxic or pharmacologic activity. Absorption is the entry of the substance into the organism, normally by uptake into the bloodstream via mucous surfaces, such as the intestinal walls of the digestive tract during ingestion. Low solubility, chemical instability in the stomach, and inability to permeate the intestinal wall can all reduce the extent to which a substance is absorbed after being ingested. The solubility of a chemical in fat and its hydrophobicity influence the pathways by which it is metabolized (structurally transformed) and whether it persists in the body or is excreted. Absorption is a critical determinant of the chemical’s bioavailability, that is, the fraction of it that reaches the systemic circulation. Other routes of absorption ex- perienced by free-ranging humans are inhalation (entry via the airways) and der- mal exposure (entry via the skin). Animal studies may involve additional routes of exposure that are not ordinarily encountered by humans, such as intravenous or intraperitoneal injection, in which the chemical is injected into the bloodstream or abdominal cavity, respectively. Distribution refers to the travel of a substance from the site of entry to the tissues and organs where they will have their ultimate effect or be sequestered. Distribution takes place most commonly via the bloodstream. The term metabo- lism is used to describe the breaking down that all substances begin to experience as soon as they enter the body. Metabolism of most foreign substances takes place in the liver by the action of oxidative enzymes collectively termed cyto- chrome P450. As metabolism occurs, the initial (parent) chemical is converted to new chemicals called metabolites. When metabolites are pharmacologically or toxicologically inert, metabolism deactivates the administered dose of the parent chemical reducing its effects on the body. Sometimes metabolism may activate the compound to a metabolite more potent or more toxic than the parent compound. Excretion, also referred to as elimination, is the removal of substances or

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 67 their metabolites from the body, most commonly in urine or feces. Excretion is often incomplete, and incomplete excretion results in the accumulation of foreign substances that can adversely affect function. The routes and rates of absorption, distribution, metabolism, and excretion of a toxic substance collectively are termed toxicokinetics (or pharmacokinetics). Those processes determine the amount of a particular substance or metabolite that reaches specific organs or cells and that persists in the body. Understanding the toxicokinetics of a chemical is important for valid reconstruction of exposure of humans and for assessing the risk of effects of a chemical. The principles involved in toxicokinetics are similar among chemicals, although the degree to which different processes influence the distribution depends on the structure and other inherent properties of the chemicals. Thus, the lipophilicity or hydrophobicity of a chemical and its structure influence the pathways by which it is metabolized and whether it persists in the body or is excreted. The degree to which different toxicokinetic processes influence the toxic potential of a chemical depends on metabolic pathways, which often differ among species. For that reason, attempts at extrapolation from experimental animal studies to human exposures must be done with extreme care. Many chemicals were used by the US armed forces in Vietnam. The nature of the substances themselves was discussed in more detail in Chapter 6 of the original VAO report (IOM, 1994). Four herbicides documented in military records were of particular concern and are examined here: 2,4-D, 2,4,5-T, 4-amino-3,5,6- trichloropicolinic acid (picloram), and cacodylic acid (dimethyl arsenic acid, DMA). This chapter also examines 2,3,7,8-tetrachlorodibenzo-p-dioxin (referred to in this report as TCDD to represent a single, and the most toxic, congener of the tetrachlorodibenzo-p-dioxins [tetraCDDs], also commonly referred to as dioxin), a contaminant of 2,4,5-T, because its potential toxicity is of concern; considerably more information is available on TCDD than on the herbicides. Other contaminants present in 2,4-D and 2,4,5-T are of less concern. Except as noted, the laboratory studies of the chemicals of concern used pure compounds or formulations; the epidemiologic studies discussed in later chapters often tracked exposures to mixtures. TCDD Chemistry TCDD is a polychlorinated dibenzo-p-dioxin that has a triple-ring structure consisting of two benzene rings connected by an oxygenated ring (Figure 4-1); chlorine atoms are attached at the 2, 3, 7, and 8 positions of the benzene rings. The chemical properties of TCDD include a molecular weight of 322, a melting point of 305–306°C, a boiling point of 445.5°C, and a log octanol–water parti- tion coefficient of 6.8 (NTP substance profile). It is virtually insoluble in water

68 VETERANS AND AGENT ORANGE: UPDATE 2008 Cl O Cl Cl O Cl 2,3,7,8-tetrachlorodibenzo-p-dioxin FIGURE 4-1  Chemical structure of TCDD. Figure 4-1.eps (19.3 ng/L), but is soluble in organic solvents, such as benzene and acetone. It has been suggested (EPA 2004 Draft Document) that volatilization of dioxin from water may be an important mechanism of transfer from the aqueous to the atmospheric phase. Absorption, Distribution, Metabolism, and Elimination The absorption, distribution, metabolism, and elimination of TCDD have been extensively studied in a number of animal models in the last 25 years. Given the plethora of data, this section only highlights and summarizes key find- ings. A more exhaustive review may be found at http://www.epa.gov/ncea/pdfs/ dioxin/nas-review. TCDD is absorbed into the body rapidly but is eliminated slowly. Because of the slow elimination, the concentration of TCDD in lipid or blood is thought to be in dynamic equilibrium with that in other tissue compartments and is thus considered to be reasonable for use in estimating total body burdens. Exposure of humans to TCDD is thought to occur primarily via the mouth, skin, and lungs. In laboratory animals, oral administration of TCDD has been shown to result in absorption of 50–93% of the administered dose (Nolan et al., 1979; Rose et al., 1976). Similarly, a study performed in a 42-year-old man found that 87% of the oral dose was absorbed. Dermal absorption appears to be dose-dependent, with lower absorption occurring at higher doses (Banks and Birnbaum, 1991). Studies performed in humans indicate that human skin may be more resistant to absorp- tion (Weber, 1991). After ingestion and gastrointestinal absorption, TCDD associates primarily with the lipoprotein fraction of the blood and later partitions into the cellular membranes and tissues (Henderson and Patterson, 1988). TCDD is distributed to all compartments of the body; the amounts differ from organ to organ, but most studies indicate that the primary disposition of TCDD is in the liver and adipose tissues. For example, in a human volunteer, it was found that at 135 days after ingestion, 90% of TCDD was in fat (Poiger and Schlatter, 1986); in the rhesus monkey, TCDD is very persistent in adipose tissue (Bowman et al., 1989). The

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 69 disposition and elimination of TCDD depend on the tissue examined, the time that has elapsed since exposure, total exposure, and other factors. For example, the concentration of cytochrome P450 1A2 (CYP1A2) (Poland et al., 1989) in the liver is increased by TCDD. Direct binding of TCDD to the CYP1A2 is thought to result in sequestration of TCDD in the liver and to inhibit its distribution to other tissues. The importance of CYP1A2 concentrations for the toxic actions of TCDD has also been shown in studies performed in laboratory animals in which maternal hepatic CYP1A2 was found to sequester TCDD and protect the fetus against TCDD-induced teratogenesis (Dragin et al., 2006). In addition, distribu- tion of TCDD is age-dependent, as shown by studies in which young animals displayed the highest concentration of TCDD in the liver and older animals the highest concentrations in kidney, skin, and muscle (Pegram et al., 1995). Finally, the elimination rate of TCDD, in particular after low exposures, depends heavily on the amount of adipose tissue mass (Aylward et al., 2005; Emond et al., 2005, 2006). In laboratory animals and humans, metabolism of TCDD occurs slowly. It is eliminated primarily in feces as both the parent compound and its more polar metabolites. However, elimination appears to be dose-dependent; at low doses, about 35% of the administered dose of TCDD was detected in the feces; at higher doses, about 46% was observed (Diliberto et al., 2001). The dose-dependent occurrence of TCDD metabolites in the feces is thought to be due to increased expression of metabolizing enzymes at higher doses. A measure of elimination is half-life, which is defined as the time required for the plasma concentration or the amount of a chemical in the body to be reduced by one-half. The half-life of TCDD in humans varies with body mass index, age, sex, and concentration and has been found to vary from 0.4 to over 10 years (Table 4-1). In light of the variables discussed above and the effect of differences in physiologic states and metabolic processes, which can affect the mobilization of lipids and possibly of compounds stored in them, complex models known as physiologically based pharmacokinetic models have been developed to integrate exposure dose with organ mass, blood flow, metabolism, and lipid content to predict the movement of toxicants into and out of each organ. A number of recent modeling studies have been performed in an effort to understand the relevance of animal experimental studies to exposures that occur in human populations (Aylward et al., 2005a,b; Emond et al., 2005). Toxicity Profile The administration of TCDD to laboratory animals affects many tissues and organs.  The effects of TCDD in laboratory animals have been observed in a number of species (rats, mice, guinea pigs, hamsters, monkeys, cows, and rabbits) after the administration of a variety of doses and after periods that represent acute (less than 24 hr), subchronic (more than 1 day up to 3 months), and chronic (more

70 VETERANS AND AGENT ORANGE: UPDATE 2008 TABLE 4-1  Estimates of TCDD Half-Life in Humans and Animals Confidence Reference Half-Lifea Interval Comment Human studies: Leung et al., 2006 0.4 year Breast-fed infants, 0–1 year after exposure Kumagai and Koda, 2005 1.1–2.3 years Adult male, incinerator workers, 0–1.3 years after exposure Aylward et al., 2005a < 3 years Calculated for exposures > 10,000 pg/g of serum lipid > 10 years Calculated for exposures < 50 pg/g of serum lipid Flesch-Janys et al., 1996 7.2 years Adult males, Boehringer cohort Geusau et al., 2002 1.5 yearsb Adult female, severe exposure, 0–3 years after exposure 2.9 yearsb Adult female, severe exposure, 0–3 years after exposure Michalek et al., 2002 0.34 yearb Adult males, Seveso cohort, 0–3 months after exposure 6.9 years Adult males, Seveso cohort, 3–16 years after exposure 9.8 years Adult females, Seveso cohort, 3–16 years after exposure 7.5 years Adult males, Ranch Hands, 9–33 years after exposure Needham et al., 1994 7.8 years 7.2–9.7 Adults, Seveso cohort years Pirkle et al., 1989 7.1 years 5.8–9.6 Adult males, Ranch Hands, 9–23 years years after exposure Animal studies: Neubert et al., 1990 73.7 days 60.9–93.8 Monkeys, marmoset, single injection days DeVito and Birnbaum, 1995 15 days Mice, female B6C3F1 Gasiewicz et al., 1983 11.0 daysc Mice, C5BL/6J 24.4 daysc Mice, DBA/2J 12.6 daysc Mice, B6D2F1/J Koshakji et al., 1984 20 days Mice, male ICR/Ha Swiss Hurst et al., 1998 8 days Rats, Long-Evans, excretion from liver Pohjanvirta and Tuomisto, 21.9 days Rats, male Han/Wistar, resistant 1990 strain Viluksela et al., 1996 20.2 days Rats, Long-Evans, TurkuAB strain 28.9 daysd Rats, Long-Evans, Charles River strain Weber et al., 1993 16.3 ± 3.0 days Rats, male Sprague-Dawley a Half-livesof TCDD in humans based on measurement of TCDD in serum samples. b Shorter half-lives measured in humans during first months after exposure or in severely contami- nated persons consistent with nonlinear elimination predicted by physiologically based pharmacoki- netic modeling (for example, by Carrier et al., 1995). Greater half-life in females attributed to greater body-mass index. c Total cumulative excretion of 3H-TCDD-derived radioactivity.

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 71 than 3 months) exposures. Some differences are observed in the different species, particularly with respect to their degree of sensitivity, but in general the effects observed are qualitatively similar. Relatively high exposures of TCDD affect a variety of organs and result in organ dysfunction and death. The specific organ dysfunction that constitutes the lethal event, however, is not known. A character- istic of TCDD exposure is a wasting syndrome with loss of adipose and muscle tissues and severe weight loss. In most rodents, exposure to TCDD affects the liver, as indicated by hepatic enlargement, the presence of hepatic lesions, and impaired hepatic function. The thymus is also sensitive. Finally, in both humans and nonhuman primates, TCDD exposure results in chloracne and associated dermatologic changes. As will be discussed in more detail in Chapters 6–9, stud- ies performed in animal models have indicated that exposure to TCDD adversely affects the heart, the skin, and the immune, endocrine, and reproductive systems, and increases the incidence of cancers of the liver, skin, thyroid, adrenal cortex, hard palate, nasal turbinates, tongue, and respiratory and lymphatic systems (Huff et al., 1994). When TCDD has been administered to pregnant animals, such birth defects as cleft palate, malformations of the reproductive organs of the male and female progeny, and abnormalities in the cardiovascular system have been observed. The administration of TCDD to laboratory animals and cultured cells affects enzymes, hormones, and receptors.  In addition to adversely affecting the ability of specific organs to fulfill their normal physiologic roles, TCDD has been found to alter the function and expression of essential proteins. Some of the proteins are enzymes, specialized proteins that increase the rates of chemical reactions and aid in the body’s ability to convert chemicals into different molecules. The metabolism of foreign chemicals often changes their biologic properties and in some cases increases the body’s ability to eliminate them in urine. The enzymes that are most affected by TCDD are ones that act on or metabolize xenobiotics and hormones. Xenobiotics are chemicals that are not expected to be present in the body, and hormones are made by the body and serve as chemical messengers that transport a signal from one cell to another. Among the enzymes affected by TCDD, the best studied is CYP1A1, which metabolizes xenobiotics. In labora- tory animals, exposure to TCDD commonly results in an increase in the CYP1A1 present in most tissues; CYP1A1 therefore is often used as a marker of TCDD exposure. Other enzymes that are affected by TCDD are ones that metabolize hormones such as thyroid hormones, retinoic acid, testosterone, estrogens, and adrenal ste- roids. Those hormones transmit their signals by interacting with specific proteins called receptors and in this manner initiate a chain of events in many tissues of the body. For example, binding of the primary female sex hormone, estrogen, to the estrogen receptor promotes the formation of breasts and the thickening of the endometrium and regulates the menstrual cycle. Exposure to TCDD can increase

72 VETERANS AND AGENT ORANGE: UPDATE 2008 the metabolism of estrogen, and this leads to a decrease in the amount of estrogen available for binding and activating the estrogen receptor. The ultimate effect of TCDD is an interference with all the bodily functions that are regulated by estro- gens. Similarly, the actions of TCDD on the adrenal steroids can adversely affect their ability to regulate glucose tolerance, insulin sensitivity, lipid metabolism, obesity, vascular function, and cardiac remodeling. In addition to changing the amount of hormone present, TCDD has been found to interfere with the ability of receptors to fulfill their role in transmitting hormone signals. Animal models have shown that exposure to TCDD can increase the amounts of enzymes in the body and interfere with the ability of hormones to activate their specific hormone re- ceptors. Those actions of TCDD on enzymes and hormone receptors are thought to underlie, in part, observed developmental and reproductive effects and cancers that are hormone-responsive. TCDD alters the paths of cellular differentiation.  Research performed primarily in cultured cells has shown that TCDD can affect the ability of cells to undergo such processes as proliferation, differentiation, and apoptosis. During the pro- liferative process, cells grow and divide. When cells are differentiating, they are undergoing a change from less specialized to more specialized. Cellular dif- ferentiation is essential for an organism to mature from a fetal to an adult state. In the adult, proper differentiation is required for normal functions of the body, for example, in maintaining a normally responsive immune system. Processes of controlled cell death, such as apoptosis, are similarly important during develop- ment of the fetus and are necessary for normal physiologic functions in the adult. Apoptosis is a way for the body to eliminate damaged or unnecessary cells. The ability of a cell to undergo proliferation, differentiation, and apoptosis is tightly controlled by an intricate network of signaling molecules that allows the body to maintain the appropriate size and number of all the specialized cells that form the fabric of complex tissues and organs. Disruption of that network that alters the delicate balance of cell fate can have severe consequences, including impair- ment of the function of the organ because of the absence of specialized cells. Alternatively, the presence of an excess of some kinds of cells can result in the formation and development of tumors. Thus, the ability of TCDD to disrupt the normal course of a specific cell to proliferate, differentiate, or undergo apoptosis is thought to underlie (at least in part) its adverse effects on the immune system and the developing fetus and its ability to promote the formation of certain cancers. Definition of Dioxin-like Compounds and TEF and TEQ Terminology Many compounds have dioxin-like properties: they have similar chemical structure, have similar physiochemical properties, and cause a common battery of toxic responses. Because of their hydrophobic nature and resistance to me-

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 73 tabolism, these chemicals persist and bioaccumulate in fatty tissues of animals and humans. Several hundred chemicals—such as the polychlorinated dibenzo- p-dioxins, polychlorinated dibenzofurans, polybrominated dibenzo-p-dioxins, polybrominated dibenzofurans, and polychlorinated biphenyls—are described as dioxin-like compounds (DLCs), although only a few of them are thought to display dioxin-like toxicity. For most purposes, only 17 polychlorinated dibenzo- p-dioxins and polychlorinated dibenxofurans and a few of the coplanar poly- chlorinated biphenyls that are often encountered in environmental samples are recognized as being true DLCs. In the context of risk assessment, these polychlo- rinated dibenzo-p-dioxins, polychlorinated dibenxofurans, and polychlorinated biphenyls are commonly found as complex mixtures when detected in environ- mental media and biologic tissues or when measured as environmental releases from specific sources. That complicates the human health risk assessment that may be associated with exposures to varied mixtures of DLCs. To address the problem, the concept of toxic equivalency has been elaborated by the scientific community, and the toxic equivalency factor (TEF) has been developed and in- troduced to facilitate risk assessment of exposure to those chemical mixtures. On the most basic level, TEFs compare the potential toxicity of each DLC found in a mixture with the toxicity of TCDD, the most toxic member of the group. The procedure involves assigning individual TEFs to the DLCs with consideration of chemical structure, persistence, and resistance to metabolism. TEF ascribe spe- cific order-of-magnitude toxicity to each DLC relative to that of TCDD, which is assigned a TEF of 1.0. The DLCs have TEFs ranging from 0.00001 to 1.0. When several compounds are present in a mixture, the toxicity of the mixture is estimated by multiplying the TEF of each DLC in the mixture by its mass concentration and summing the products to yield the TCDD toxicity equivalent quotient (TEQ) of the mixture. Mechanism of Action TCDD binds and activates the aryl hydrocarbon receptor (AHR). The AHR is a member of a family of basic-helix-loop-helix (bHLH) transcription factors, that is, one of many proteins in the cell that controls the transfer (or transcription) of genetic information from DNA to RNA. bHLH proteins are characterized by the presence of a string of basic amino acid residues followed by two alpha helices joined by a loop. Generally, the larger of the two helices participates in binding to DNA in a specific sequence motif; the specificity is determined largely by the amino acid sequence of the helix. bHLH transcription factors are dimeric, form- ing functional heterodimers with other members of the family. By mechanisms that are poorly understood, binding of the heterodimeric complex to DNA recruits the transcriptional machinery needed to activate gene expression and results in a large increase in the rate of synthesis of mRNA molecules for the genes regulated by the complex and ultimately in a large increase in the corresponding protein.

74 VETERANS AND AGENT ORANGE: UPDATE 2008 The best known AHR target is the expression of a mixed-function oxidase en- zyme that was termed aryl hydrocarbon hydroxylase (AHH) in the 1960s and is now better known as the CYP1A1 enzyme. Expression of this protein is an acute outcome of AHR activation and may not faithfully represent the consequences of chronic exposure to AHR ligands. In its inactive state, the AHR is found in the cytosol of the cell, where it is protected from proteolytic degradation by several chaperones and cochaperones. As a receptor, the AHR is a protein capable of receiving and forming a com- plex with specific substances, termed ligands, which confer on it the ability to perform a biologic function. In the case of the AHR, the function is to induce the transcription of specific target genes. Hence, the AHR belongs to a class of ligand-activated transcription factors. If the ligand is a chemical, such as TCDD, the AHR dissociates from the chaperones and translocates into the nucleus of the cell, where it forms a heterodimer with another bHLH protein, the AHR nuclear translocator (ARNT). This heterodimer binds to its cognate DNA motifs and recruits the macromolecular complexes needed to initiate gene transcription. AHR Functional Domains The AHR contains several regions, or domains, that perform distinct func- tions. The receptor is a member of the Per-Arnt-Sim (PAS) bHLH subfamily (Burbach et al., 1992; Fukunaga et al., 1995). The bHLH motif is found in the amino terminus of the protein and is common to all transcription factors in this subfamily (Jones, 2004). The members of the bHLH family also have several highly conserved domains with functionally distinctive biochemical roles. One of the domains is the basic region, described earlier, which is involved in the binding of AHR/ARNT complexes to DNA. Another domain is the HLH region, which facilitates the stable interaction between AHR and ARNT. A third domain is termed the PAS domain and consists of a stretch of 200–350 amino acids with high sequence relatedness to protein domains that were originally found in the Drosophila melanogaster genes period (Per) and single minded (Sim) and in the AHR’s dimerization partner ARNT; hence the name PAS. The AHR contains two PAS domains, PAS-A and PAS-B (Ema et al., 1992). The PAS domains support secondary interactions with other PAS-domain–containing proteins, with the chaperones and cochaperones, and with many other transcription factors, coacti- vators and corepressors. The ligand-binding site of AHR is in the PAS-B domain (Coumailleau et al., 1995) and contains several conserved residues critical for ligand binding (Goryo et al., 2007). A fourth important domain in the carboxyl terminus of the protein, is rich in glutamine and is involved in coregulator recruit- ment and transactivation (Kumar et al., 2001).

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 75 AHR Ligands From an environmental point of view, there are two classes of AHR ligands— synthetic and naturally occurring—that total more than 400 known ligands. Many of the first ligands to be discovered were synthetic polycyclic aromatic hydrocar- bons (PAHs), such as 3-methylcholanthrene, benzo[a]pyrene (B[a]P), benzanthra- cene, and naphthoflavone. The biologic consequences of experimental exposure of mice to those chemicals led to the prediction of a receptor-dependent mecha- nism long before the existence of the AHR was directly demonstrated. Compari- son of the effects of 3-methylcholanthrene treatment in two inbred mouse strains revealed a major difference in PAH responsiveness. Hepatic CYP1A1 enzyme increased more than 6-fold after 3-methylcholanthrene treatment in C57BL/6 mice but not in DB/2 mice. Appropriate genetic crosses between responsive C57BL/6 mice but in non-responsive DB/2 mice indicated that responsiveness in these prototype strains was inherited as a simple autosomal dominant trait. The genetic locus defined in the crosses was termed the aromatic hydrocarbon responsiveness (Ahr) locus (Nebert et al., 1982). Molecular biologic studies dur- ing the next decade showed that responsive and nonresponsive mice had equally functional CYP1A1 enzymes and that the Ahr locus encoded a regulatory gene responsible for induction of the Cyp1a1 gene. The members of the polyhalo- genated aromatic hydrocarbons—such as the dibenzodioxins, dibenzofurans, and polychlorinated and polybrominated biphenyls—were recognized as AHR ligands much later, after the discovery by Poland and co-workers that dioxin was a potent inducer of hepatic AHH in the rat. At that time, it was found that high concentrations of TCDD could induce AHH activity in the nonresponsive DB/2 mouse to levels as high as those in the responsive C57BL/6 mouse. The difference between responsive and nonresponsive strains was in sensitivity to the inducer: DB/2 mice required 18 times more TCDD than C57BL/6 mice for 50% of the maximal response. Later, a receptor protein for TCDD, 3-methylcholanthrene and other PAHs, was identified, characterized in the hepatic cytosol of C57BL/6 mice, and termed Ah receptor (Poland et al., 1976). The available evidence indicated that the protein was the product of the Ahr locus, which was localized to mouse chromosome 12 and human chromosome 7, and later cloned (Burbach et al., 1992; Ema et al., 1992). Recent work has focused on naturally occurring compounds in hopes of identifying an endogenous ligand (Denison and Nagy, 2003). Several such natu- rally occurring compounds have been identified as AHR ligands, including the tryptophan derivatives indigo and indirubin (Adachhi et al., 2001), the tetrapyr- roles bilirubin (Sinal and Bend, 1997), the arachidonic acid metabolites lipoxin A4 and prostaglandin G (Seidel et al., 2001), modified low-density lipoprotein (McMillan and Bradfield, 2007), several dietary carotenoids (Denison and Negy, 2003; Probst et al., 1993), and cAMP (Oesch-Bartlomowicz et al., 2005). One assumption made in the search for an endogenous ligand is that the ligand will

76 VETERANS AND AGENT ORANGE: UPDATE 2008 be a receptor agonist, that is, that it will activate the receptor. However, recent work has shown that may not be the case inasmuch as one such natural ligand, 7-ketocholesterol, competitively inhibits AHR-dependent signal transduction (Savouret et al., 2001). AHR Signaling Pathway In the absence of bound ligand, the inactive AHR is retained in the cyto- plasm of the cell in a complex consisting of two molecules of the heat shock protein hsp90, one molecule of prostaglandin E synthase 3 (p23) (Kazlauskas et al., 1999), and one molecule of the immunophilin-like protein hepatitis B virus X-associated protein 2 (XAP2) (Petrulis et al., 2003), previously identified as AHR-interacting protein (Ma and Whitlock, 1997) and AHR-activated 9 (Carver and Bradfield, 1997). The hsp90 dimer–p23 complex plays multiple roles in the protection of the AHR from proteolysis, maintaining it in a conformation that makes the receptor accessible to ligand binding at the same time that it prevents the premature binding of ARNT (Carver and Bradfield, 1994; Pongratz et al., 1992; Whitelaw et al., 1993). XAP2 interacts with the carboxyl terminus of hsp90 and with the AHR nuclear-localization signal (NLS), a short amino acid domain in the bHLH region that targets the receptor for interaction with nuclear-transport proteins. Binding of XAP2 blocks such interaction, preventing the inappropriate trafficking of the receptor into the nucleus (Petrulis et al., 2003). Binding of ligand induces the release of XAP2 and the exposure of the NLS and leads to the binding of nuclear-import proteins and translocation of the cytosolic complex into the nucleus (Davarinos and Pollenz, 1999; Song and Pollenz, 2002). Once in the nucleus, chaperones and cochaperones dissociate from the AHR, exposing the two PAS domains and allowing the binding of ARNT (Hoffman et al., 1991; Probst et al., 1993). The activated AHR/ARNT heterodimeric complex is then capable of directly or indirectly interacting with DNA by binding to recognition sequences in the regulatory region of responsive genes (Dolwick et al., 1993; Probst et al., 1993). The canonical DNA recognition motif of the AHR/ARNT complex is re- ferred to as the AHR-responsive element (AHRE, also referred to as the dioxin- responsive element [DRE] or the xenobiotic-responsive element [XRE]). This element is found in the promoter region of AHR-responsive genes and contains the core sequence 5′-GCGTG-3′ (Shen and Whitlock, 1992), which is part of a more extensive consensus binding sequence, 5′-T/GNGCGTGA/CG/CA-3′ (Lusska et al., 1993; Yao and Denison, 1992). The AHR/ARNT complex binds to the AHRE core sequence in such a manner that ARNT binds to the 5′-GTG-3′ and AHR binds to 5′-TC/TGC-3′ (Bacsi et al., 1995; Swanson et al., 1995). A second type of element, termed AHRE-II, 5′-CATG(N6)C[T/A]TG-3′, has recently been shown to be capable of acting indirectly with the AHR/ARNT complex (Boutros et al., 2004; Sogawa et al., 2004). The end result of the process is the recruitment

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 77 of the transcriptional machinery associated with RNA polymerase II and the initiation of differential changes in the expression of the genes bearing the AHR/ ARNT recognition motif. Many of the genes code for proteins responsible for detoxification reactions directed at the elimination of the ligand. Recent research suggests that posttranslational modifications in histone proteins may modify the response (Hestermann and Brown, 2003; Schnekenburger et al., 2007). AHR Physiology The vertebrate AHR is presumed to have evolved from its counterpart in invertebrates, in which it serves a ligand-independent role in normal develop- ment processes. The ancestral function of the AHR appears to be the regulation of specific aspects of embryonic development, it having acquired the ability to bind xenobiotic compounds only during vertebrate evolution (Hahn, 2001). The invertebrate AHR also functions as a transcription factor and binds to the same dimerization partner (ARNT) and DNA response elements as the vertebrate pro- tein, but it does not respond to any of the environmental ligands recognized by the vertebrate receptor. Instead, it regulates diverse developmental processes that are independent of exogenous ligand exposure, such as neuronal differentiation dur- ing worm development in Caenorhabditis elegans (Huang et al., 2004; Qin and Powell-Coffman, 2004) or normal morphogenesis of legs, antennae, and bristles in Drosophila melanogaster (Adachi-Yamada et al., 2005). In developing verte- brates, the AHR seems to play a role in cellular proliferation and differentiation and, in keeping with this role in invertebrates, also possesses a developmental role in craniofacial, renal, and cardiovascular morphogenesis (Birnbaum et al., 1989; Fernandez-Salguero et al., 1997; Lahvis et al., 2005). The clearest adaptive physiologic response of AHR activation is the induc- tion of xenobiotic-metabolizing enzymes involved in detoxification of toxic li- gands. Evidence of that response, which was described above, was first observed from the induction of Cyp1a1, resulting from exposure to PAHs or TCDD, and directly related to activation of the AHR signaling pathway (Israel and Whitlock, 1983, 1984). Because of the presence of AHRE motif in their gene promoters, other metabolizing genes were tested and found to be induced by AHR ligands, and this led to the identification of a so-called AHR gene battery of phase I and phase II detoxification genes that code for the drug-metabolizing enzymes CYP1A1, CYP1A2, CYP1B1, NQO1, ALHD3A1, UGT1A2, and GSTA1 (Ne- bert et al., 2000). Presumably, vertebrates have evolved those enzymes to detect a wide array of foreign, potentially toxic chemicals, represented in the wide variety of substrates that AHR is able to bind whose biotransformation and elimination it is able to facilitate. A potential complication of the adaptive responses elicited by AHR activa- tion is the induction of a toxic response. Toxicity may result from the adaptive response itself if the induction of metabolizing enzymes results in the production

78 VETERANS AND AGENT ORANGE: UPDATE 2008 of toxic metabolites. For example, the PAH B[a]P, an AHR ligand, induces its own metabolism and detoxification by the AHR-dependent signaling mechanism described earlier but paradoxically becomes bioactivated to a toxic metabolite in several tissues by metabolism that depends on CYP1A1 and CYP1B1 enzymatic activity (Harrigan et al., 2004). A second potential source of AHR-mediated tox- icity may be aberrant changes in global gene expression beyond those observed in the AHR gene battery. The global changes in gene expression may lead to deleterious changes in cellular processes and physiology. Microarray analysis has proved invaluable in understanding and characterizing that response (Martinez et al., 2002; Puga et al., 2000, 2004; Vezina et al., 2004). It is clear that the AHR is an essential component of the toxicity of dioxin and of DLCs. Homozygous deletion of the AHR in mice leads to a phenotype that is resistant to the toxic effects of TCDD and to the carcinogenic effects of B[a]P (Fernandez-Salguero et al., 1996; Lahvis and Bradfield, 1998; Schmidt et al., 1996). AHR knockout mice, however, have other phenotypic effects, including reduced liver size and hepatic fibrosis and cardiovascular abnormalities. Hence, it is likely that dioxin has effects due to gratuitous deregulation of endogenous AHR functions, unrelated to the intrinsic toxicity of some of its ligands. Carcinogenic Classification of TCDD The US Environmental Protection Agency (EPA) and the International Agency for Research on Cancer (IARC), a branch of the World Health Organiza- tion, have defined criteria to classify the potential carcinogenicity of chemicals on the basis of the weight of scientific evidence from animal, human, epidemiologic, mechanistic, and mode-of-action studies. EPA classified TCDD as a “probable human carcinogen” in 1985 and as “carcinogenic to humans” in a 2003 reassess- ment. In 1998, the IARC panel of experts concluded that the weight of scientific evidence supported the classification of dioxin as a class I carcinogen, that is, as “carcinogenic to humans.” Four years later, the US National Toxicology Program upgraded its classification to “known to be a human carcinogen.” In 2006, a panel of experts convened by the National Research Council to evaluate the EPA reas- sessment concluded that TCDD was “likely to be carcinogenic to humans.” That designation reflected the revised EPA Guidelines for Carcinogen Risk Assessment made public in 2005. Other Toxic Health Outcomes of Dioxin There is an extensive body of evidence from experimental studies in animal model systems that TCDD, other dioxins, and several DLCs are immunotoxic. Although the available evidence on humans is scant, mechanistic considerations support the notion that chemical alterations to immune function would cause adverse health outcomes because of the critical role that the immune system

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 79 plays in general protection, fighting off infection and eliminating cancer cells at early stages. Because of those considerations, these compounds are potential immunotoxicants. Similarly, reproduction and embryonic development clearly are targets of TCDD, other dioxins, and DLCs; it is found consistently that the adverse effects are more prevalent during fetal development than in the adult. However, data on those effects in humans are practically nonexistent. TCDD, other dioxins, and DLCs are also recognized as potentially capable of causing birth defects, reproductive disorders, immunotoxicity, and chloracne. Human and animal studies have revealed other potential health outcomes includ- ing cardiovascular disease, hepatic disease, thyroid dysfunction, lipid disorders, neurotoxicity, and metabolic disorders, such as diabetes. A number of effects of TCDD exposure in vitro appear to be independent of AHR-mediated mechanisms. The salient ones are induction of transforming growth factor-α and other genes involved in extracellular matrix deposition in cells from mice with homozygous ablation of the Ahr gene (Guo et al., 2004); mobilization of calcium from intracellular sources and imported from the culture medium (Puga et al., 1995); and related to calcium mobilization, the induction of mitochondria oxidative stress (Senft et al., 2002). Calcium mobilization by TCDD may have an important effect on signal-transduction mechanisms that control gene expression, inasmuch as several proto-oncogenes, such as c-fos, are activated by calcium changes. Recent Findings from Mechanistic Studies That May Be Relevant to Health Outcomes Several reports published since Update 2006 (IOM, 2007) have contributed to the understanding of how TCDD exposure may have contributed to particular health outcomes, as follows: • Insight into the role of the AHR repressor protein in mediating cell- type–specific responsiveness to TCDD (Evans et al., 2008; Haarmann- Stemmann et al., 2007). • Identification of genes that may be regulated by TCDD through activation of AHR that are involved in cholesterol biosynthesis, lipogenesis, and glu- cose metabolism (Sato et al., 2008; Zarzour et al., 2008); development of thymocytes (the KLF2 regulon) (McMillian et al., 2007); glucose uptake (the glucose transporter) (Tonack et al., 2007); and metabolism of the ste- roid hydroxysteroid 17-beta dehydrogenase; plus a novel gene (Hectd2) with poorly defined function (Hayes et al., 2007). • Examination of the effect of TCDD on signaling pathways, such as that of protein kinase C (Lee et al., 2007) and integrin signaling (Liu and

80 VETERANS AND AGENT ORANGE: UPDATE 2008 Jefcoate, 2006) that may underlie TCDD’s adverse effects on neurons in cell culture and cell adhesion, respectively. • Elucidation of crosstalk between the AHR and receptors involved in regulation of cellular responses to stress (Burchiel et al., 2008; Dvorak et al., 2008; Sonneveld et al., 2007) and estrogen (Boverhof et al., 2008; Marquez-Bravo and Gierthy, 2008; Tanaka et al., 2007). • Progress pertaining to TCDD’s effects on oxidative stress and DNA dam- age (Lin et al., 2007; Lu et al., 2007) and events in the mitochondria (Biswas et al., 2008; Shertzer et al., 2006). • Study of TCDD’s effects on determination of cell-fate. Mechanisms by which TCDD induces G1 arrest in hepatic cells (Mitchell et al., 2006; Weiss et al., 2008) and decreases viability of endometrial endothelial cells (Bredhult et al., 2007), insulin-secreting beta cells (Piaggi et al., 2007), peripheral T cells (Singh et al., 2008), and neuronal cells (Bredhult et al., 2007) were identified. The findings may be relevant to cancer, reproduc- tive health, diabetes, immune function, and neurotoxicity. Insights were also gained into TCDD’s effects on maturation and differentiation of adipocytes (Arsenescu et al., 2008), granule neuron precursors (Collins et al., 2008), dendritic cells (Lee et al., 2007), and how TCDD induces the differentiation of regulatory T cells (Funatake et al., 2008; Gill et al., 2008; Hollingshead et al., 2008; Kimura et al., 2008; Quintana et al., 2008; Veldhoen et al., 2008; Vogel et al., 2008). • Increased evidence that the TCDD–AHR pathway may impinge on the cytokine-inflammatory response (Chiaro et al., 2008; Dong and Matsumura, 2008; Hollingshead et al., 2008; Ito et al., 2008; Li et al., 2007; Vogel et al., 2007a,b). Those publications collectively have added to what was already strong sup- port of an association of TCDD exposure with adverse health effects in laboratory animals and, by extension, in humans. Summary on Biologic Plausibility of TCDD Inducing Adverse Effects in Humans Mechanistic studies in vitro and in laboratory animals have characterized the biochemical pathways and types of biologic events that contribute to adverse effects of exposure to TCDD. For example, much evidence indicates that TCDD acting via the AHR in partnership with ARNT alters gene expression. Receptor binding may result in release of other cytoplasmic proteins that alter the expres- sion or activity of other cell-regulatory proteins. Mechanistic studies also indicate that many other cellular component proteins contribute to the gene-regulatoring effect and that the response to TCDD exposure involves a complex interplay between genetic and environmental factors. Comparative data from animal and

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 81 human cells in vitro and from tissues suggest a strong qualitative similarity among species in response to TCDD, and this further supports the applicability to humans of the generalized model of initial events in response to dioxin ex- posure. Biochemical and biologic responses to TCDD exposure are considered adaptive or simply reflective of exposure and not adverse in themselves if they take place within the normal homeostatic parameters of an organism. However, they may exceed physiologic parameters or constitute an early event in a pathway leading to damage to sensitive members of the population. In the latter case, the response is toxic and would be expected to cause an adverse health effect. From a mechanistic standpoint, adverse effects identified in vitro are expected to occur in all the cells of an organism and to all the organisms that express these proteins. That generalization sets the ground rules for the concept of biologic plausibil- ity, which relies on extrapolation from laboratory tests to human risks, and on the precautionary principle, which bases decision-making on precaution if the precise nature or magnitude of the potential damage that a substance may cause in humans is uncertain. When deemed useful, the findings from individual publications are presented in the biologic-plausibility sections associated with specific health outcomes. PHENOXY HERBICIDES: 2,4-D AND 2,4,5-T Chemistry 2,4-D (Chemical Abstracts Service [CAS] No. 94-75-7) is an odorless and, when pure, white crystalline powder (Figure 4-2); it may appear yellow when phenolic impurities are present. The melting point of 2,4-D is 138°C, and the free acid is corrosive to metals. It is soluble in water and in a variety of organic sol- vents (such as acetone, alcohols, ketones, ether, and toluene). 2,4,5-T (CAS No. 93-76-5) is an odorless, white to light-tan solid with a melting point of 158°C. FIGURE 4-2  Structure of selected phenoxy herbicides. Figure 4-2.eps

82 VETERANS AND AGENT ORANGE: UPDATE 2008 2,4,5-T is noncorrosive and is soluble in alcohol and water. It reacts with organic and inorganic bases to form salts and with alcohols to form esters. Uses of 2,4-D and 2,4,5-T 2,4-D has been used commercially in the United States since World War II to control the growth of broadleaf plants and weeds on range lands, lawns, golf courses, forests, roadways, parks, and agricultural land and remains today a widely used herbicide approved for use by the European Union and the US EPA. Formulations include 2,4-D amine and alkali salts and esters, which are mobile in soil and easily absorbed through the leaves and roots of many plants. Like 2,4-D, 2,4,5-T was developed and marketed as a herbicide during World War II. How- ever, the registration for 2,4,5-T was canceled by EPA in 1978 when it became clear that it was contaminated with TCDD during the manufacturing process. The herbicidal properties of 2,4-D and 2,4,5-T are related to their ability to mimic the plant growth hormone indole acetic acid. They are selective herbicides in that they affect the growth of only broadleaf dicots (which include most weeds) and do not affect monocots, such as wheat, corn, and rice. Absorption, Distribution, Metabolism, and Elimination Several studies have examined the absorption, distribution, metabolism, and excretion of 2,4-D and 2,4,5-T in animals and humans. Data on both compounds are consistent among species and support the conclusion that absorption of oral or inhaled doses is rapid and complete. Absorption through the skin is much lower but may be increased with the use of sunscreens or alcohol (Brand et al., 2002; Pont et al., 2004). After absorption, 2,4-D and 2,4,5-T are distributed widely in the body but are eliminated quickly, predominantly in unmetabolized form in urine (Sauerhoff et al., 1977). Neither 2,4-D nor 2,4,5-T is metabolized to a great extent in the body although 2,4,5-trichlorophenol and 2,4-dichlorophenol have been identified as trace metabolites in urine. The half-life in humans after single doses of 2,4-D or 2,4,5-T has been estimated to be about 18–23 hours (Gehring et al., 1973; Kohli et al., 1974; Sauerhoff et al., 1977; WHO, 1984). Results of a recent study that examined concentrations of 2,4-D and its metabolites in the urine of herbicide applicators was consistent with 2,4-D urinary half-life esti- mates of 13–40 hours in humans (Hines et al., 2003). Toxicity Profile The toxicity data base on 2,4-D is extensive (http://www.epa.gov/ttn/atw/ hlthef/di-oxyac.html: accessed January 21, 2009), whereas the data available on the toxicity of purified 2,4,5-T, independent of its contamination by TCDD, are sparse. TCDD is much more toxic than 2,4,5-T, and much of the toxicity at-

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 83 tributed to 2,4,5-T in early studies was later shown to be caused by the TCDD contaminant. The following summary therefore focuses on 2,4-D toxicity, and information on pure 2,4,5-T is added when it is available. After a single oral dose, 2,4-D is considered to produce moderate acute toxic- ity with an LD50 (dose lethal to 50% of exposed animals) of 375 mg/kg in rats, 370 mg/kg in mice, and from less than 320 to 1,000 mg/kg in guinea pigs. Rats and rabbits have dermal LD50s of 1,500 mg/kg and 1,400 mg/kg, respectively. 2,4,5-T itself also produces moderate acute toxicity, with oral LD50s of 389 mg/kg in mice and 500 mg/kg in rats. Death from acute poisoning with 2,4-D or 2,4,5-T has been attributed to the ability of the chemicals to uncouple oxidative phosphorylation, a vital process used by almost all cells in the body as the pri- mary means of generating energy. After exposure to high doses, death can occur rapidly from multiple organ failure. Studies in rats, cats, and dogs indicate that the central nervous system is the principal target organ for acute 2,4-D toxicity in mammals and suggest that the primary site of action is the cerebral cortex or the reticular formation (Arnold et al., 1991; Dési et al., 1962a,b). Neurotoxicity in humans is the predominant effect of acute inhalation and oral exposure to 2,4-D; symptoms include stiffness of arms and legs, incoordination, lethargy, anorexia, stupor, and coma. 2,4-D is also an irritant of the gastrointestinal tract, causing nausea, vomiting, and diarrhea. Chronic exposure to 2,4-D at relatively high concentrations has been shown to produce a variety of toxic effects, including hepatic and renal toxicity, neuro- toxicity, and hematologic changes. A no-observed-effect level (NOEL) of 2,4-D of 1 mg/kg was identified for renal toxicity in rats (Hazleton Laboratories Amer- ica, 1987). The reproductive toxicity of 2,4-D is limited to reduced survival and decreased growth rates of offspring of mothers fed high doses during pregnancy and was associated with maternal toxicity. However, even at high exposures, 2,4- D did not affect fertility and did not produce teratogenic effects in the offspring. The purity of 2,4,5-T has been shown to influence its reproductive toxicity; TCDD contamination increases its fetotoxic effects and induces teratogenic ef- fects. Immunotoxicity of 2,4-D has been reported in a small number of studies. At high doses that produced clinical toxicity, suppression of the antibody response was observed, whereas other measures of immune function were normal. The immunotoxicity of 2,4,5-T has not been evaluated in laboratory animals. Carcinogenicity The carcinogenicity of 2,4-D or 2,4,5-T has been studied in rats, mice, and dogs after exposure in their food, direct placement in their stomachs, or expo- sure of their skin. All the studies had negative results except one that found an increased incidence of brain tumors in male rats, but not female rats, that received the highest dose of 2,4-D. The occurrence of malignant lymphoma in dogs kept as pets was reported to be higher when owners reported that they used 2,4-D on their

84 VETERANS AND AGENT ORANGE: UPDATE 2008 lawns than when they did not (Hayes et al., 1991, 1995), but detailed reanalysis did not confirm this finding (Kaneene and Miller, 1999). A controlled study using dogs exposed to 2,4-D in the laboratory had negative results. Timchalk (2004) suggested that dogs are not relevant for comparative evaluation of human health risk attributable to 2,4-D exposure, because they excrete 2,4-D less efficiently than rats or humans. 2,4-D is not metabolized to reactive intermediates capable of interacting with DNA, and the evidence supports the conclusion that 2,4-D is not a carcinogen. CACODYLIC ACID Chemistry Arsenic (As) is a naturally occurring element that exists in trivalent form (As+3 or AsIII) and pentavalent form (As+5 or AsV). The AsIII in sodium arsenite is generally considered to be the most toxic—see Figure 4-3 for chemical struc- FIGURE 4-3  Structure of selected arsenic-containing compounds. Figure 4-3.eps

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 85 tures of selected arsenic-containing compounds. Arsenic is commonly present in drinking-water sources associated with volcanic soils and can reach high concentrations (over 50 ppb). Numerous human health effects have been attrib- uted to drinking-water exposure, particularly bladder, skin, and lung cancers and vascular diseases. Arsenic exists in both inorganic and organic (methylated) forms and is read- ily metabolized in humans and other species. Inorganic arsenic can be converted to organic forms, but organic forms cannot be converted into inorganic forms (Cohen et al., 2006). Cacodylic acid has a valence of +5 and is commonly referred to as dimethylarsinic acid (DMAV). Cacodylic acid, disodium methanearsonate, and monosodium methanearsonate are herbicides that EPA-approved for use in the United States, where they are occasionally applied on golf courses and large open spaces. Cacodylic acid was the form of arsenic used in Agent Blue; DMAV made up about 30% of Agent Blue, one of the mixtures used for defoliation in Vietnam. Agent Blue was chemically and toxicologically unrelated to Agent Or- ange, which consisted of phenoxy herbicides contaminated with DLCs. As shown in Figure 4-4, DMA and monomethyl arsonic acid (MMA) are metabolic products of exposure to inorganic arsenic. Methylation of inorganic arsenic used to be considered a detoxification process associated with increased excretion (Vahter and Concha, 2001). More recently, however, some of the methylated metabolic intermediates have been thought to be more toxic than the parent compound. The MMAIII iAsV Limited MMAV cellular uptake DMAV DMAs 60–80% of human urinary excretion iAsIII iAsIII DMAIII Extensive TMA cellular uptake TMAO TMAO MMAs None found in 10–20% of human human urine, urinary excretion 5–10% of rat urinary excretion FIGURE 4-4  General pathways of arsenic metabolism after exposure to inorganic arsenic (iAs). SOURCE: Adapted with permission from Cohen et al., 2006. Figure 4-4.eps

86 VETERANS AND AGENT ORANGE: UPDATE 2008 methylation pathway of inorganic arsenic results in the formation of pentavalent DMA (DMAV) and trivalent DMA (DMAIII). The committee contemplated the relevance to DMA of data on exposure to inorganic arsenic. Although inorganic arsenic is a human carcinogen, there is no evidence that direct exposure to DMA produces cancer in humans. DMA also is not demethylated to inorganic arsenic (Vahter et al., 1984). It has not been established, nor can it be inferred, that the observed effects of exposure to inorganic arsenic are also caused by exposure to DMA. Therefore, the literature on inorganic arsenic is not considered in this report. The reader is referred to Arsenic in Drinking Water (NRC, 1999a) and Arsenic in Drinking Water: 2001 Update (NRC, 2001). Toxicokinetics The metabolism and disposition of DMAV has recently been reviewed (Cohen et al., 2006). In general, DMAV is rapidly excreted mostly unchanged in the urine of most animal species after systemic exposure. However, rats are unique in that a small percentage (10%) of DMAV binds to hemoglobin in red blood cells and leads to a longer half-life in blood (Cui et al., 2004; Suzuki et al., 2004). Rat he- moglobin is unique in its binding of DMAV in that its binding affinity to DMAV is 10 times higher than that of human hemoglobin (Lu et al., 2004). Chronic expo- sure of normal rat hepatocytes to DMAV resulted in reduced uptake over time and in acquired cytotoxic tolerance (Kojima et al., 2006); the tolerance was mediated by induction of glutathione-S-transferase activity and of multiple-drug–resistant protein expression. Adair et al. (2007) recently examined the tissue distribution of DMA in rats after dietary exposures for 14 days; they found that it was extensively metabolized to trimethylated forms that may play a role in toxicity. Recently, a physiologically based pharmacokinetic model for intravenous and ingested DMAV has been developed based on mouse data (Evans et al., 2008). Similar models have been developed for humans on the basis of exposure to inor- ganic arsenic (El-Masri and Kenyon, 2008), but these models have limited utility in considering the toxicity of DMAV exposures that are relevant to veterans. Toxicity Profile This section discusses the toxicity associated with organic forms of arsenic, most notably DMAV because this is the active ingredient in Agent Blue. The toxicity of inorganic arsenic is not considered relevant to veteran exposures to Agent Blue. Neurotoxicity Kruger et al. (2006) found that DMAIII and DMAV significantly attenu- ated neuronal ion currents through N-methyl-D-aspartate receptor ion channels

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 87 whereas only DMAV inhibited ion currents through α-amino-3-hydroxy-5- methylisoxazole-4-propionic acid receptors. The data suggest that those methyl- ated forms of arsenic may have neurotoxic potential. Immunotoxicity Previous studies have shown that a low concentration of DMAV (10–7 M) could increase proliferation of human peripheral blood monocytes after their stim- ulation with phytohemagglutinin whereas it took a high concentration (10 –4 M) to inhibit release of interferon-γ; this suggested that immunomodulatory effects of DMAV are concentration-specific (Di Giampaolo et al., 2004). Genotoxicity and Carcinogenicity Cancer has been induced in the urinary bladder, kidneys, liver, thyroid glands, and lungs of laboratory animals exposed to high concentrations of DMA (Wei et al., 2002). Exposure to DMAV resulted in necrosis of the urinary bladder epithelium followed by regenerative hyperplasia (Cohen et al., 2002). DMAIII was considerably more potent than DMAV in inducing DNA dam- age in Chinese hamster ovary cells (Dopp et al., 2004), and this was associated with a 10% uptake of DMAIII into the cells compared with 0.03% uptake of DMAV. An additional study showed that DMAV is poorly membrane-permeable, but when forced into cells by electroporation it can induce DNA damage (Dopp et al., 2005). Furthermore, DMAV induced protein-DNA adducts in lung fi- broblast cells (MRC-5) (Mouron et al., 2005) and transformation loci in 3T3 fibroblasts following subsequent treatment with the tumor promoter 12-O- tetradecanoylphorbol-13-acetate (TPA) (Tsuchiya et al., 2005). However, DMAV was devoid of promotion activity in 3T3 fibroblasts when cells were pretreated with 3-methylcholanthrene or sodium arsenite. DMAV, but not inorganic arsenic species, exhibited genotoxicity in Drosophila as assessed with the somatic mutation and recombination test (Rizki et al., 2006). Drosophila lacks the ability to methylate arsenic, so the data suggest that arsenic biomethylation is a key determinant of arsenic genotoxicity. DMAIII and DMAV have been shown to induce DNA damage by increasing oxidative stress. Chronic exposure of ddY mice to DMAV at 400 ppm in drinking water increased staining for 4-hydroxy-2-nonenal adducts, which are indicative of oxidative stress, and for 8-oxo-2′-deoxyguanosine (8-oxodG), reactive oxygen- species–induced DNA damage, in Clara cells of the lung (An et al., 2005). Gomez et al. (2005) demonstrated that DMAIII induced a dose-related increase in DNA damage and oxidative stress in Jurkat cells. Two studies investigated the degree to which oxidative stress may mediate DMAV cytotoxicity. In one study, an antioxidant (N-acetylcysteine, vitamin C, or melatonin) and DMAV at 100 ppm were coadministered to F344 rats for 10 weeks

88 VETERANS AND AGENT ORANGE: UPDATE 2008 (Wei et al., 2005). N-Acetylcysteine inhibited DMAV-induced proliferation of the urinary bladder epithelium whereas neither vitamin C nor melatonin had an ef- fect; this suggested that oxidative stress may mediate the cytotoxic process in the urothelium. In the second, metallothionein wild-type and null mice were exposed to a single oral dose of DMAV at 0, 188, 375, or 750 mg/kg (Jia et al., 2004). DMAV induced a dose-dependent increase in metallothionein in the livers of wild-type mice, but metallothionein was undetectable and uninducible in the null mice. At 24 hours after exposure, DMAV induced dose-dependent DNA adducts, DNA strand breaks, and pulmonary and bladder apoptosis in both genotypes, but the incidence of damage was significantly higher in the null mice. Those results suggest that metallothionein may play a protective role against DMAV-induced DNA damage. Gene-expression profiling of bladder urothelium after chronic exposure to DMAV in drinking water showed significant increases in genes that regulate apop- tosis, the cell cycle, and oxidative stress (Sen et al., 2005). Furthermore, doses that were nontoxic, according to a lack of histologic and ultrastructural changes, could be distinguished from toxic doses on the basis of the expression of a subset of genes involved in control of cell signaling and the stress response, such as thioredoxin E-cadherin. Xie et al. (2004) administered DMAV at 1,000 ppm in drinking water to v-Ha-ras transgenic mice for 17 weeks and after 4 weeks of treatment applied TPA to the skin twice a week. The results were an initial 10% body-weight loss, a cumulative mortality of 20%, hepatic arsenic accumulation, hepatocellular degeneration and foci of inflammation without evidence of hepatic tumors, and hepatic DNA hypomethylation. Hepatic gene-expression profiling showed that DMAV exposure induced changes consistent with oxidative stress, including induction of heme oxygenase, NAD(P)H:quinone oxidoreductase, and glutathione-S-transferase. Mizoi et al. (2005) found that chronic administration of DMAV at 400 ppm to mice after their initiation with 4-nitroquinolone 1-oxide (4NQO) significantly increased the number of lung tumors and the percentage of mice that had lung tumors. DMAV also significantly increased pulmonary 8-oxodG adducts regard- less of whether the mice had been treated with 4NQO. Hairless mice treated with DMAIII on the skin after initiation with dimethylbenz[a]anthracene exhibited a significant increase in epidermal 8-oxodG adducts and skin tumors. In a 2-year bioassay, F344 rats were exposed to DMAV at 0, 2, 10, 40, or 100 ppm in drinking water, and C57BL/6 mice were exposed at 0, 8, 40, 200, or 500 ppm (Arnold et al., 2006). The rats developed epithelial carcinomas and papillomas in the urinary bladder and nonneoplastic changes in the kidneys. The mice failed to develop any tumors but exhibited glomerular nephropathy, nephrocalcinosis, and vacuolation of the urinary epithelium. The murine NOEL based on nonneoplastic changes was 40 ppm in males and 8 ppm in females; the rat NOEL based on neoplastic and nonneoplastic changes was 10 ppm in both sexes. In a recent study, Cohen et al. (2007) exposed F344 rats to DMAV at

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 89 2–100 ppm in the diet for 2 years and found an increase in bladder tumors; they postulated that trimethylated forms of arsenic may be responsible for bladder cancer in rats. Similar findings have not been reported in other species. In light of the significant differences in metabolism of arsenic by different species and the lack of supportive data in humans, it cannot be concluded that DMAV leads to an increase in cancer risk in humans. Mechanisms Oxidative stress is a common theme that runs through the literature on the mechanisms of action of arsenic, particularly with regard to cancer in animals, although some studies have suggested that methylated arsenicals (MMAIII and DMAIII) can induce mutations in mammalian cells at concentrations below those required to produce oxidative stress after in vitro exposures (Klein et al., 2008). Recent studies have shown that mice deficient in DNA-repair enzymes associated with oxidative stress are highly susceptible to formation of tumors, particularly lung tumors, induced by DMAV (Kinoshita et al., 2007). The chemical reaction of arsenicals with thiol groups in sensitive target tissues, such as red blood cells and kidneys, may also be a mechanism of action of organic arsenicals (Naranmandura and Suzuki, 2008). The variation in the susceptibility of various animal species to tumor for- mation caused by inorganic and organic arsenic is thought to depend heavily on differences in metabolism and distribution. Thus, genetic differences may play an important role. Numerous investigators are examining potential human susceptibility factors and gene polymorphisms that may increase a person’s risk of cancer and other diseases induced by arsenicals. Several such studies were reported during the update period covered by this update (Hernandez et al., 2008; Huang SK et al., 2008; Huang YK et al., 2008; McCarty et al., 2007; Meza et al., 2007; Steinmaus et al., 2007). However, the studies are in early stages, and it is not possible to identify polymorphisms that may contribute to a person’s suscep- tibility to DMA-induced cancer or tissue injury. PICLORAM Chemistry Picloram (4-amino-3,5,6-trichloropyridine-2-carboxylic acid or 4-amino- 3,5,6-trichloropicolinic acid; see chemical structure in Figure 4-5) was used with 2,4-D in the herbicide formulation Agent White, which was sprayed in Vietnam. Picloram is also used commonly in Australia in a formulation with the trade name Tordon 75D®. Tordon 75D contains several chemicals, including 2,4-D, picloram, a surfactant diethyleneglycolmonoethyl ether, and a silicone defoamer. A number

90 VETERANS AND AGENT ORANGE: UPDATE 2008 Picloram [1918-02-1] CI NH 2 HO CI O N CI FIGURE 4-5  Structure of picloram. Figure 4-5.eps of studies of picloram used such mixtures as Tordon or other mixtures of 2,4-D and picloram that are similar to Agent White. Toxicokinetics The original VAO committee reviewed studies of the toxicokinetics of pi- cloram. Studies of animals showed rapid absorption through the gastrointestinal tract and rapid elimination of picloram as the unaltered parent compound in urine. Nolan et al. (1984) examined the toxicokinetics of picloram in six healthy male volunteers who were given single oral doses of 0.5 or 5.0 mg/kg and a dermal dose of 2.0 mg/kg. Picloram was rapidly absorbed in the gavage study and rapidly excreted as unchanged compound in urine. More than 75% of the dose was excreted within 6 hours, and the remainder with an average half-life of 27 hours. On the basis of the quantity of picloram excreted in urine in the skin study, the authors noted that only 0.2% of the picloram applied to the skin was absorbed. Because of its rapid excretion, picloram has low potential to accumu- late in humans. In general, the literature on picloram toxicity continues to be sparse. Studies of humans and animals indicate that picloram is rapidly eliminated as the parent compound. Studies of animals have indicated that picloram is sparingly toxic at high doses. Toxicity Profile The original VAO committee reviewed studies of the carcinogenicity, geno- toxicity, acute toxicity, chronic systemic toxicity, reproductive and developmental toxicity, and immunotoxicity of picloram. In general, there is limited evidence for cancer in some rodent models but not in other species (NCI, 1978). In those studies, there was some concern that contaminants in the picloram (in particular, hexachlorobenzene) might be responsible for the carcinogenicity. Therefore, picloram has not been established as a chemical carcinogen. There is also no evidence that picloram is a genotoxic agent on the basis

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 91 of studies conducted by EPA (1988c). Picloram is considered a mild irritant; erythema is seen in rabbits only at high doses. The available information on the acute toxicity of picloram is also paltry. Some neurologic effects—including hyperactivity, ataxia, and tremors—were reported in pregnant rats exposed to picloram at 750 or 1,000 mg/kg (Thompson et al., 1972). Chronic Systemic Toxicity Several studies have reported various effects of technical-grade picloram on the livers of rats. In the carcinogenicity bioassay conducted by Stott and col- leagues (1990) described above, treatment-related hepatomegaly, hepatocellular swelling, and altered tinctorial properties in the central regions of the liver lobules were noted in the groups exposed at 60 and 200 mg/kg per day. In addition, males and females exposed at the high dose had higher liver weights than controls. The NOEL was 20 mg/kg per day, and the lowest effect level was 60 mg/kg per day for histologic changes in centrilobular hepatocellular tissues. According to EPA, hexachlorobenzene (at 197 ppm) was probably not responsible for the hepatic ef- fects (EPA, 1988c). Gorzinski and colleagues (1987) also reported a dose-related increase in liver weights, hepatocellular hypertrophy, and changes in centrilobular tinctorial properties in male and female F344 rats exposed to picloram at 150 mg/kg per day and higher in the diet for 13 weeks. In a 90-day study, cloudy swelling in the liver cells and bile duct epithelium occurred in male and female F344 rats given 0.3% or 1% technical picloram in the diet (EPA, 1988c). Hepatic effects have also been reported in dogs exposed to picloram: increased liver weights were reported in beagles that received 35 mg/kg per day or more in the diet for 6 months (EPA, 1988c). No other effects of chronic exposure to picloram have been reported. Reproductive and Developmental Toxicity The reproductive toxicity of picloram was evaluated in a two-generation study; however, too few animals were evaluated, and no toxicity was detected at the highest dose tested, 150 mg/kg per day (EPA, 1988c). Some developmental toxicity was produced in rabbits exposed to picloram by gavage at 400 mg/kg per day on days 6–18 of gestation. Fetal abnormalities included single-litter inci- dences of forelimb flexure, fused ribs, hypoplastic tail, and omphalocele (John- Greene et al., 1985). Some maternal toxicity was observed at that dose, however, and EPA concluded on the basis of the low-litter incidence of the findings, that the malformations were not treatment-related (EPA, 1988c). No teratogenic effects were produced in the offspring of rats given picloram by gavage at up to 1,000 mg/kg per day on days 6–15 of gestation, although the occurrence of bilateral accessory ribs was significantly increased (Thompson et al., 1972).

92 VETERANS AND AGENT ORANGE: UPDATE 2008 Immunotoxicity Studies of the potential immunotoxicity of picloram included dermal sen- sitization and rodent immunoassays. In one study, 53 volunteers received nine 24-hour applications of 0.5 mL of a 2% potassium picloram solution on the skin of both upper arms. Each volunteer received challenge doses 17–24 days later. The formulation of picloram (its potassium salt) was not a skin sensitizer or an irritant (EPA, 1988c). In a similar study, a 5% solution of picloram (M-2439, Tordon 101 formulation) produced slight dermal irritation and a sensitization response in 6 of the 69 volunteers exposed. When the individual components of M-2439—picloram, triisopropanolamine (TIPA) salt, and 2,4-D TIPA salt—were tested separately, no sensitization reaction occurred (EPA, 1988c). Tordon K+, but not technical-grade picloram, was also found to be a skin sensitizer in guinea pigs (EPA, 1988c). CD1 mice exposed to Tordon 202C (94% 2,4-D and 6% piclo- ram) had no consistent adverse effects on antibody responses (Blakley, 1997). Mechanisms There are no well-characterized mechanisms of toxicity known for picloram, and therefore they are not discussed here. REFERENCES Adachi J, Mori Y, Matsui S, Takigami H, Fujino J, Kitagawa H, Miller C III, Kato T, Saeki K, Matsuda T. 2001. Indirubin and indigo are potent aryl hydrocarbon receptor ligands present in human urine. Journal of Biological Chemistry 276:31475–31478. Adachi-Yamada T, Harumoto T, Sakurai K, Ueda R, Saigo K, O’Connor MB, Nakato H. 2005. Wing- to-leg homeosis by spineless causes apoptosis regulated by fish-lips, a novel leucine-rich repeat transmembrane protein. Molecular & Cellular Biology 25(8):3140–3150. Adair BM, Moore T, Conklin SD, Creed JT, Wolf DC, Thomas DJ. 2007. Tissue distribution and uri- nary excretion of dimethylated arsenic and its metabolites in dimethylarsinic acid- or arsenate- treated rats. Toxicology and Applied Pharmacology 222(2):235–242. An Y, Kato K, Nakano M, Otsu H, Okada S, Yamanaka K. 2005. Specific induction of oxidative stress in terminal bronchiolar Clara cells during dimethylarsenic-induced lung tumor promoting process in mice. Cancer Letters 230(1):57–64. Arnold EK, Beasley VR, Parker AJ, Stedelin JR. 1991. 2,4-D toxicosis II: A pilot study of clinical pathologic and electroencephalographic effects and residues of 2,4-D in orally dosed dogs. Veterinary and Human Toxicology 33:446–449. Arnold LL, Eldan M, Nyska A, van Gemert M, Cohen SM. 2006. Dimethylarsinic acid: Results of chronic toxicity/oncogenicity studies in F344 rats and in B6C3F1 mice. Toxicology 223(1-2): 82–100.   Throughout the report the same alphabetic indicator following year of publication is used con- sistently for the same article when there were multiple citations by the same first author in a given year. The convention of assigning the alphabetic indicator in order of citation in a given chapter is not followed.

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 93 Arsenescu V, Arsenescu RI, King V, Swanson H, Cassis LA. 2008. Polychlorinated biphenyl–77 induces adipocyte differentiation and proinflammatory adipokines and promotes obesity and atherosclerosis. Environmental Health Perspectives 116(6):761–768. Aylward LL, Brunet RC, Carrier G, Hays SM, Cushing CA, Needham LL, Patterson DG Jr, Gerthoux PM, Brambilla P, Mocarelli P. 2005a. Concentration-dependent TCDD elimination kinetics in humans: Toxicokinetic modeling for moderately to highly exposed adults from Seveso, Italy, and Vienna, Austria, and impact on dose estimates for the NIOSH cohort. Journal of Exposure Analysis and Environment Epidemiology 15(1):51–65. Aylward LL, Brunet RC, Starr TB, Carrier G, Delzell E, Cheng H, Beall C. 2005b. Exposure recon- struction for the TCDD-exposed NIOSH cohort using a concentration- and age-dependent model of elimination. Risk Analysis 25(4):945–956. Bacsi SG, Reisz-Porszasz S, Hankinson O. 1995. Orientation of the heterodimeric aryl hydrocarbon (dioxin) receptor complex on its asymmetric DNA recognition sequence. Molecular Pharma- cology 47(3):432–438. Banks YB, Birnbaum LS. 1991. Absorption of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) after low dose dermal exposure. Toxicology and Applied Pharmacology 107(2):302–310. Birnbaum L, Harris M, Stocking L, Clark A, Morrissey R. 1989. Retinoic acid and 2,3,7,8-tetrachloro­ dibenzo-p-dioxin selectively enhance teratogenesis in C57BL/6N mice. Toxicology and Applied Pharmacology 98:487–500. Biswas G, Srinivasan S, Anandatheerthavarada HK, Avadhani NG. 2008. Dioxin-mediated tumor progression through activation of mitochondria-to-nucleus stress signaling. Proceedings of the National Academy of Sciences of the United States of America 105(1):186–191. Blakley BR. 1997. Effect of Roundup and Tordon 202C herbicides on antibody production in mice. Veterinary and Human Toxicology 39(4):204–206. Boutros PC, Moffat ID, Franc MA, Tijet N, Tuomisto J, Pohjanvirta R, Okey AB. 2004. Dioxin-re- sponsive AHRE–II gene battery: Identification by phylogenetic footprinting. Biochemical and Biophysical Research Communications 321(3):707–715. Boverhof DR, Burgoon LD, Williams KJ, Zacharewski TR. 2008. Inhibition of estrogen-mediated uterine gene expression responses by dioxin. Molecular Pharmacology 73(1):82–93. Bowman RE, Schantz SL, Weerasinghe NCA, Gross ML, Barsotti DA. 1989. Chronic dietary intake of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) at 5 and 25 parts per trillion in the monkey: TCDD kinetics and dose–effect estimate of reproductive toxicity. Chemosphere 18:243–252. Brand RM, Spalding M, Mueller C. 2002. Sunscreens can increase dermal penetration of 2,4- dichlorophenoxyacetic acid. Journal of Toxicology–Clinical Toxicology 40(7):827–832. Bredhult C, Backlin BM, Olovsson M. 2007. Effects of some endocrine disruptors on the prolif- eration and viability of human endometrial endothelial cells in vitro. Reproductive Toxicology 23(4):550–559. Burbach KM, Poland A, Bradfield CA. 1992. Cloning of the Ah–receptor cDNA reveals a distinctive ligand-activated transcription factor. Proceedings of the National Academy of Sciences of the United States of America 89:8185–8189. Burchiel SW, Thompson TA, Lauer FT, Oprea TI. 2008. Corrigendum to “Activation of dioxin re- sponse element (DRE)–associated genes by benzo(a)pyrene 3,6-quinone and benzo(a)pyrene 1,6-quinone in MCF-10A human mammary epithelial cells.” Toxicology and Applied Pharma- cology 226(3):345–346. Carrier G, Brunet RC, Brodeur J. 1995. Modeling of the toxicokinetics of polychlorinated dibenzo- p-dioxins and dibenzofurans in mammalians, including humans. II. Kinetics of absorption and disposition of PCDDs/PCDFs. Toxicology and Applied Pharmacology 131(2):267–276. Carver L, Bradfield C. 1997. Ligand-dependent interaction of the aryl hydrocarbon receptor with a novel immunophilin homolog in vivo. Journal of Biological Chemistry 272:11452–11456. Carver L, Jackiw V, Bradfield C. 1994. The 90-kDa heat shock protein is essential for Ah receptor signaling in a yeast expression system. Journal of Biological Chemistry 269:30109–30112.

94 VETERANS AND AGENT ORANGE: UPDATE 2008 Chiaro CR, Morales JL, Prabhu KS, Perdew GH. 2008. Leukotriene A4 metabolites are endogenous ligands for the Ah receptor. Biochemistry 47(32):8445–8455. Cohen SM, Arnold LL, Uzvolgyi E, Cano M, St John M, Yamamoto S, Lu X, Le XC. 2002. Possible role of dimethylarsinous acid in dimethylarsinic acid–induced urothelial toxicity and regenera- tion in the rat. Chemical Research in Toxicology 15(9):1150–1157. Cohen SM, Arnold LL, Eldan M, Lewis AS, Beck BD. 2006. Methylated arsenicals: The implications of metabolism and carcinogenicity studies in rodents to human risk assessment. Critical Reviews in Toxicology 36(2):99–133. Cohen SM, Ohnishi T, Arnold LL, Le XC. 2007. Arsenic-induced bladder cancer in an animal model. Toxicology and Applied Pharmacology 222(3):258–263. Collins LL, Williamson MA, Thompson BD, Dever DP, Gasiewicz TA, Opanashuk LA. 2008. 2,3,7,8- tetracholorodibenzo-p-dioxin exposure disrupts granule neuron precursor maturation in the developing mouse cerebellum. Toxicological Sciences 103(1):125–136. Coumailleau P, Poellinger L, Gustafsson J, Whitelaw M. 1995. Definition of a minimal domain of the dioxin receptor that is associated with hsp90 and maintains wild type ligand binding affinity and specificity. Journal of Biological Chemistry 270:25291–25300. Cui X, Kobayashi Y, Hayakawa T, Hirano S. 2004. Arsenic speciation in bile and urine following oral and intravenous exposure to inorganic and organic arsenics in rats. Toxicological Sciences 82(2):478–487. Davarinos N, Pollenz R. 1999. Aryl hydrocarbon receptor imported into the nucleus following ligand binding is rapidly degraded via the cytosplasmic proteasome following nuclear export. Journal of Biological Chemistry 274:28708–28715. Denison M, Nagy S. 2003. Activation of the aryl hydrocarbon receptor by structurally diverse exog- enous and endogenous chemicals. Annual Review of Pharmacology Toxicology 43:309–334. Dési I, Sos J, Nikolits I. 1962a. New evidence concerning the nervous site of action of a chemical herbicide causing professional intoxication. Acta Physiologica Academiae Scientiarum Hun- garicae 22:73–80. Dési I, Sos J, Olasz J, Sule F, Markus V. 1962b. Nervous system effects of a chemical herbicide. Archives of Environmental Health 4:95–102. DeVito M, Birnbaum L. 1995. The importance of pharmacokinetics in determining the relative po- tency of 2,3,7,8-tetrachlorodibenzo-p-dioxin and 2,3,7,8-tetrachlorodibenzofuran. Fundamental and Applied Toxicology 24:145–148. Di Giampaolo L, Di Gioacchino M, Qiao N, Travaglini P, D’Intino A, Kouri M, Ponti J, Castellani ML, Reale M, Gabriele E, Boscolo P. 2004. “In vitro” effects of different arsenic compounds on PBMC (preliminary study). Giornale Italiano di Medicina del Lavoro Ed Ergonomia 26(3): 183–186. Diliberto JJ, Devito MJ, Ross DG, Birnbaum LS. 2001. Subchronic exposure of [3H]-2,3,7,8- tetrachlorodibenzo-p-dioxin (TCDD) in female B6C3F1 mice: Relationship of steady-state levels to disposition and metabolism. Toxicological Sciences 61:241–255. Dolwick KM, Swanson HI, Bradfield CA. 1993. In vitro analysis of Ah receptor domains involved in ligand-activated DNA recognition. Proceedings of the National Academy of Sciences (USA) 90:8566–8570. Dong B, Matsumura F. 2008. Roles of cytosolic phospholipase A2 and Src kinase in the early ac- tion of 2,3,7,8-tetrachlorodibenzo-p-dioxin through a nongenomic pathway in MCF10A cells. Molecular Pharmacology 74(1):255–263. Dopp E, Hartmann LM, Florea AM, von Recklinghausen U, Pieper R, Shokouhi B, Rettenmeier AW, Hirner AV, Obe G. 2004. Uptake of inorganic and organic derivatives of arsenic associated with induced cytotoxic and genotoxic effects in Chinese hamster ovary (CHO) cells. Toxicology and Applied Pharmacology 201(2):156–165.

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 95 Dopp E, Hartmann LM, von Recklinghausen U, Florea AM, Rabieh S, Zimmermann U, Shokouhi B, Yadav S, Hirner AV, Rettenmeier AW. 2005. Forced uptake of trivalent and pentavalent methylated and inorganic arsenic and its cyto-/genotoxicity in fibroblasts and hepatoma cells. Toxicological Sciences 87(1):46–56. Dragin N, Dalton TP, Miller ML, Shertzer HG, Nebert DW. 2006. For dioxin-induced birth defects, mouse or human CYP1A2 in maternal liver protects whereas mouse CYP1A1 and CYP1B1 are inconsequential. Journal of Biological Chemistry 281(27):18591–18600. Dvorak Z, Vrzal R, Pavek P, Ulrichova J. 2008. An evidence for regulatory cross-talk between aryl hydrocarbon receptor and glucocorticoid receptor in HepG2 cells. Physiological Research 57(3):427–435. El-Masri HA, Kenyon EM. 2008. Development of a human physiologically based pharmacokinetic (PBPK) model for inorganic arsenic and its mono- and di-methylated metabolites. Journal of Pharmacokinetics and Pharmacodynamics 35(1):31–68. Ema M, Sogawa K, Watanabe N, Chujoh Y, Matsushita N, Gotoh O, Funae Y, Fujii-Kuriyama Y. 1992. cDNA cloning and structure of mouse putative Ah receptor. Biochemical and Biophysical Research Communications 184:246–253. Emond C, Michalek JE, Birnbaum LS, DeVito MJ. 2005. Comparison of the use of physiologically based pharmacokinetic model and a classical pharmacokinetic model for dioxin exposure as- sessments. Environmental Health Perspectives 113(12):1666–1668. Emond C, Birnbaum LS, DeVito MJ. 2006. Use of a physiologically based pharmacokinetic model for rats to study the influence of body fat mass and induction of CYP1A2 on the pharmacokinetics of TCDD. Environmental Health Perspectives 114(9):1394–1400. EPA (US Environmental Protection Agency). 1988c. Guidance for the Reregistration of Pesticide Products Containing Picloram as the Active Ingredient. Washington, DC: EPA, Office of Pes- ticide Programs. Evans BR, Karchner SI, Allan LL, Pollenz RS, Tanguay RL, Jenny MJ, Sherr DH, Hahn ME. 2008. Repression of aryl hydrocarbon receptor (AHR) signaling by AHR repressor: Role of DNA bind- ing and competition for AHR nuclear translocator. Molecular Pharmacology 73(2):387–398. Fernandez-Salguero PM, Hilbert DM, Rudikoff S, Ward JM, Gonzalez FJ. 1996. Aryl-hydrocarbon receptor-deficient mice are resistant to 2,3,7,8-tetrachlorodibenzo-p-dioxin-induced toxicity. Toxicology and Applied Pharmacology 140:173–179. Fernandez-Salguero PM, Ward JM, Sundberg JP, Gonzalez FJ. 1997. Lesions of aryl-hydrocarbon receptor-deficient mice. Veterinary Pathology 34(6):605–614. Flesch-Janys D, Becher H, Gurn P, Jung D, Konietzko J, Manz A, Papke O. 1996. Elimination of polychlorinated dibenzo-p-dioxins and dibenzofurans in occupationally exposed persons. Jour- nal of Toxicology and Environmental Health 47(4):363–378. Fukunaga BN, Probst MR, Reisz-Porszasz S, Hankinson O. 1995. Identification of functional domains of the aryl hydrocarbon receptor. Journal of Biology and Chemistry 270:29270–29278. Funatake CJ, Marshall NB, Kerkvliet NI. 2008. 2,3,7,8-tetrachlorodibenzo-p-dioxin alters the dif- ferentiation of alloreactive CD8+ T cells toward a regulatory T cell phenotype by a mechanism that is dependent on aryl hydrocarbon receptor in CD4+ T cells. Journal of Immunotoxicology 5(1):81–91. Gasiewicz T, Geiger L, Rucci G, Neal R. 1983. Distribution, excretion, and metabolism of 2,3,7,8- tetrachlorodibenzo-p-dioxin in C57BL/6J, DBA/2J, and B6D2F1/J mice. Drug Metabolism and Disposition 11:397–403. Gehring PJ, Kramer CG, Schwetz BA, Rose JQ, Rowe VK. 1973. The fate of 2,4,5-trichlorophen­ oxyacetic acid (2,4,5-T) following oral administration to man. Toxicology and Applied Phar- macology 26:352–361. Geusau A, Schmaldienst S, Derfler K, Päpke O, Abraham K. 2002. Severe 2,3,7,8-tetrachlorodibenzo- p-dioxin (TCDD) intoxication: Kinetics and trials to enhance elimination in two patients. Archives of Toxicology 76:316–325.

96 VETERANS AND AGENT ORANGE: UPDATE 2008 Gill BC, Jeon CH, Sung HN, Kim HL, Jin DW, Park JH. 2008. 2,3,7,8-tetrachlorodibenzo-p-dioxin modulates the expression of cKrox and Runx3, transcription regulatory factors controlling the lineage commitment of CD4+CD8+ into CD4 and CD8 thymocytes, respectively. Toxicology Letters 180(3):189–195. Gómez SE, Del Razo LM, Muñoz Sanchez JL. 2005. Induction of DNA damage by free radicals generated either by organic or inorganic arsenic (AsIII, MMAIII, and DMAIII) in cultures of B and T lymphocytes. Biological Trace Element Research 108(1-3):115–126. Goryo K, Suzuki A, Del Carpio CA, Siizaki K, Kuriyama E, Mikami Y, Kinoshita K, Yasumoto K, Rannug A, Miyamoto A, Fujii-Kuriyama Y, Sogawa K. 2007. Identification of amino acid residues in the Ah receptor involved in ligand binding. Biochemical and Biophysical Research Communications 354:396–402. Gorzinski SJ, Johnson KA, Campbell RA, Landry TD. 1987. Dietary toxicity of picloram herbicide in rats. Journal of Toxicology and Environmental Health 20:367–377. Guo SW. 2004. The link between exposure to dioxin and endometriosis: A critical reappraisal of primate data. Gynecologic and Obstetric Investigation 57(3):157–173. Haarmann-Stemmann T, Bothe H, Kohli A, Sydlik U, Abel J, Fritsche E. 2007. Analysis of the transcriptional regulation and molecular function of the aryl hydrocarbon receptor repressor in human cell lines. Drug Metabolism and Disposition 35(12):2262–2269. Hahn ME. 2001. Dioxin toxicology and the aryl hydrocarbon receptor: Insights from fish and other non-traditional models. Marine Biotechnology (New York, NY) 3(Supplement 1):S224–S238. Harrigan JA, Vezina CM, McGarrigle BP, Ersing N, Box HC, Maccubbin AE, Olson JR. 2004. DNA adduct formation in precision-cut rat liver and lung slices exposed to benzo[a]pyrene. Toxico- logical Sciences 77(2):307–314. Hayes HM, Tarone RE, Cantor KP, Jessen CR, McCurnin DM, Richardson RC. 1991. Case–control study of canine malignant lymphoma: Positive association with dog owner’s use of 2,4-D her- bicides. Journal of the National Cancer Institute 83(17):1226–1231. Hayes HM, Tarone RE, Cantor KP. 1995. On the association between canine malignant lymphoma and opportunity for exposure to 2,4-dichlorophenoxyacetic acid. Environmental Research 70(2): 119–125. Hayes KR, Zastrow GM, Nukaya M, Pande K, Glover E, Maufort JP, Liss AL, Liu Y, Moran SM, Vollrath AL, Bradfield CA. 2007. Hepatic transcriptional networks induced by exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin. Chemical Research in Toxicology 20(11):1573–1581. Hazleton Laboratories America. 1987. Oncogenicity Study in Mice with 2,4-Dichlorophenoxyacetic Acid (2,4-D). Final Report. Prepared for the Industry Task Force on 2,4-D Research Data. Henderson L, Patterson DJ. 1988. Distribution of 2,3,7,8-tetrachlorodibenzo-p-dioxin in human whole blood and its association with, and extractability from, lipoproteins. Bulletin of Environmental Contamination and Toxicology 40(4):604–611. Hernandez A, Xamena N, Surralles J, Sekaran C, Tokunaga H, Quinteros D, Creus A, Marcos R. 2008. Role of the Met(287)Thr polymorphism in the AS3MT gene on the metabolic arsenic profile. Mutation Research 637(1-2):80–92. Hestermann E, Brown M. 2003. Agonist and chemopreventative ligands induce differential tran- scriptional cofactor recruitment by aryl hydrocarbon receptor. Molecular and Cellular Biology 7920–7925. Hines CJ, Deddens JA, Striley CA, Biagini RE, Shoemaker DA, Brown KK, Mackenzie BA, Hull RD. 2003. Biological monitoring for selected herbicide biomarkers in the urine of exposed custom applicators: Application of mixed-effect models. Annals of Occupational Hygiene 47(6): 503–517. Hoffman EC, Reyes H, Chu FF, Sander F, Conley LH, Brooks BA, Hankinson O. 1991. Cloning of a factor required for activity of the Ah (dioxin) receptor. Science 252:954–958.

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 97 Hollingshead BD, Beischlag TV, Dinatale BC, Ramadoss P, Perdew GH. 2008. Inflammatory signal- ing and aryl hydrocarbon receptor mediate synergistic induction of interleukin 6 in MCF-7 cells. Cancer Research 68(10):3609–3617. Huang SK, Chiu AW-H, Pu Y-S, Huang Y-K, Chung C-J, Tsai H-J, Yang M-H, Chen C-J, Hsueh Y-M. 2008. Arsenic methylation capability, heme oxygenase-1 and NADPH quinone oxidore- ductase-1 genetic polymorphisms and the stage and grade of urothelial carcinomas. Urologia Internationalis 80(4):405–412. Huang X, Powell-Coffman JA, Jin Y. 2004. The AHR-1 aryl hydrocarbon receptor and its co-factor the AHA-1 aryl hydrocarbon receptor nuclear translocator specify GABAergic neuron cell fate in C. elegans. Development 131(4):819–828. Huang YK, Pu YS, Chung CJ, Shiue HS, Yang MH, Chen CJ, Hsueh YM. 2008. Plasma folate level, urinary arsenic methylation profiles, and urothelial carcinoma susceptibility. Food and Chemical Toxicology 46(3):929–938. Huff J, Lucier G, Tritscher A. 1994. Carcinogenicity of TCDD: Experimental, mechanistic, and epi- demiologic evidence. Annual Review of Pharmacology and Toxicology 34:343–372. Hurst CH, Abbott BD, DeVito MJ, Birnbaum LS. 1998. 2,3,7,8-Tetrachlorodibenzo-p-dioxin in pregnant Long Evans rats: Disposition to maternal and embryo/fetal tissues. Toxicological Sci- ences 45(2):129–136. IOM (Institute of Medicine). 1994. Veterans and Agent Orange: Health Effects of Herbicides Used in Vietnam. Washington, DC: National Academy Press. IOM. 1996. Veterans and Agent Orange: Update 1996. Washington, DC: National Academy Press. IOM. 1999. Veterans and Agent Orange: Update 1998. Washington, DC: National Academy Press. IOM. 2001. Veterans and Agent Orange: Update 2000. Washington, DC: National Academy Press. IOM. 2003. Veterans and Agent Orange: Update 2002. Washington, DC: The National Academies Press. IOM. 2005. Veterans and Agent Orange: Update 2004. Washington, DC: The National Academies Press. IOM. 2006. Health Risks from Dioxin and Related Compounds: Evaluation of the EPA Reassessment. Washington, DC: The National Academies Press. IOM. 2007. Veterans and Agent Orange: Update 2006. Washington, DC: The National Academies Press. Israel DI, Whitlock JP Jr. 1983. Induction of mRNA specific for cytochrome P1-450 in wild type and variant mouse hepatoma cells. Journal of Biological Chemistry 258:10390–10394. Israel DI, Whitlock JP Jr. 1984. Regulation of cytochrome P1-450 gene transcription by 2,3,7,8- tetrachlorodibenzo-p-dioxin in wild type and variant mouse hepatoma cells. Journal of Biologi- cal Chemistry 259:5400–5402. Ito T, Inouye K, Nohara K, Tohyama C, Fujimaki H. 2008. TCDD exposure exacerbates atopic dermatitis-related inflammation in NC/Nga mice. Toxicology Letters 177(1):31–37. Jia G, Sone H, Nishimura N, Satoh M, Tohyama C. 2004. Metallothionein (I/II) suppresses genotoxic- ity caused by dimethylarsinic acid. International Journal of Oncology 25(2):325–333. John-Greene JA, Ouellette JH, Jeffries TK, Johnson KA, Rao KS. 1985. Teratological evaluation of picloram potassium salt in rabbits. Journal of Food and Chemical Toxicology 23:753–756. Jones S. 2004. An overview of the basic helix–loop–helix proteins. Genome Biology 5:226. Kaneene JB, Miller R. 1999. Re-analysis of 2,4-D use and the occurrence of canine malignant lym- phoma. Veterinary and Human Toxicology 41(3):164–170. Kazlauskas A, Poellinger L, Pongratz I. 1999. Evidence that the Co-chaperone p23 regulates ligand responsiveness of the dioxin (Aryl hydrocarbon) receptor. Journal of Biological Chemistry 274:13519–13524. Kimura A, Naka T, Nohara K, Fujii–Kuriyama Y, Kishimoto T. 2008. Aryl hydrocarbon receptor regulates Stat1 activation and participates in the development of Th17 cells. Proceedings of the National Academy of Sciences of the United States of America 105(28):9721–9726.

98 VETERANS AND AGENT ORANGE: UPDATE 2008 Kinoshita A, Wanibuchi H, Wei M, Yunoki T, Fukushima S. 2007. Elevation of 8-hydroxydeoxy­ guanosine and cell proliferation via generation of oxidative stress by organic arsenicals contrib- utes to their carcinogenicity in the rat liver and bladder. Toxicology and Applied Pharmacology 221(3):295–305. Klein OD, Kostiner DR, Weisiger K, Moffatt E, Lindeman N, Goodman S, Tuchman M, Packman S. 2008. Acute fatal presentation of ornithine transcarbamylase deficiency in a previously healthy male. Hepatology International 2(3):390–394. Kohli JD, Khanna RN, Gupta BN, Dhar MM, Tandon JS, Sircar KP. 1974. Absorption and excretion of 2,4,5-trichlorophenoxy acetic acid in man. Archives Internationales de Pharmacodynamie et de Therapie 210:250–255. Kojima C, Qu W, Waalkes MP, Himeno S, Sakurai T. 2006. Chronic exposure to methylated arsenicals stimulates arsenic excretion pathways and induces arsenic tolerance in rat liver cells. Toxicologi- cal Sciences 91(1):70–81. Koshakji RP, Harbison RD, Bush MT. 1984. Studies on the merabolic fate of [14C] 2,3,7,8- tetrachlorodibenzo-p-dioxin (TCDD) in the mouse. Toxicology and Applied Pharmacology 73:69–77. Kruger K, Gruner J, Madeja M, Hartmann LM, Hirner AV, Binding N, Muhoff U. 2006. Blockade and enhancement of glutamate receptor responses in Xenopus oocytes by methylated arsenicals. Archives of Toxicology 80(8):492–501. Kumagai S, Koda S. 2005. Polychlorinated dibenzo-p-dioxin and dibenzofuran concentrations in serum samples of workers at an infectious waste incineration plant in Japan. Journal of Oc- cupational and Environmental Hygiene 2(2):120–125. Kumar M, Ramadoss P, Reen R, Vanden Heuvel J, Perdew G. 2001. The Q-rich sub-domain of the human Ah receptor transactivation domain is required for dioxin-mediated transcriptional activ- ity. Journal of Biology and Chemistry 276(45):42302–42310. Lahvis G, Bradfield C. 1998. Ahr null alleles: Distinctive or different? Biochemical Pharmacology 56(7):781–787. Lahvis GP, Pyzalski RW, Glover E, Pitot HC, McElwee MK, Bradfield CA. 2005. The aryl hydrocar- bon receptor is required for developmental closure of the ductus venosus in the neonatal mouse. Molecular Pharmacology 67(3):714–720. Lee JA, Hwang JA, Sung HN, Jeon CH, Gill BC, Youn HJ, Park JH. 2007. 2,3,7,8-tetrachlorodibenzo- p-dioxin modulates functional differentiation of mouse bone marrow-derived dendritic cells downregulation of RelB by 2,3,7,8-tetrachlorodibenzo-p-dioxin. Toxicology Letters 173(1): 31–40. Leung HW, Kerger BD, Paustenbach DJ. 2006. Elimination half-lives of selected polychlorinated ��������������������������������������������������� dibenzodioxins and dibenzofurans in breast-fed human infants. Journal of Toxicology and En- vironmental Health, Part A 69(6):437–443. Li W, Vogel CFA, Matsumura F. 2007. Studies on the cell treatment conditions to elicit lipolytic responses from 3T3-L1 adipocytes to TCDD, 2,3,7,8-tetrachlorodibenzo-p-dioxin. Journal of Cellular Biochemistry 102(2):389–402. Lin PH, Lin CH, Huang CC, Chuang MC, Lin P. 2007. 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) induces oxidative stress, DNA strand breaks, and poly(ADP-ribose) polymerase-1 activation in human breast carcinoma cell lines. Toxicology Letters 172(3):146–158. Liu X, Jefcoate C. 2006. 2,3,7,8-tetrachlorodibenzo-p-dioxin and epidermal growth factor coop- eratively suppress peroxisome proliferator-activated receptor-gamma1 stimulation and restore focal adhesion complexes during adipogenesis: Selective contributions of Src, Rho, and Erk distinguish these overlapping processes in C3H10T1/2 cells. Molecular Pharmacology 70(6): 1902–1915.

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 99 Lu F, Zahid M, Saeed M, Cavalieri EL, Rogan EG. 2007. Estrogen metabolism and formation of estro- gen-DNA adducts in estradiol-treated MCF-10F cells. The effects of 2,3,7,8-tetrachlorodibenzo- p-dioxin induction and catechol-O-methyltransferase inhibition. Journal of Steroid Biochemistry and Molecular Biology 105(1–5):150–158. Lu M, Wang H, Li XF, Lu X, Cullen WR, Arnold LL, Cohen SM, Le XC. 2004. Evidence of he- moglobin binding to arsenic as a basis for the accumulation of arsenic in rat blood. Chemical Research in Toxicology 17(12):1733–1742. Lusska A, Shen E, Whitlock JP Jr. 1993. Protein-DNA interactions at a dioxin-responsive enhancer. Journal of Biological Chemistry 268:6575–6580. Ma Q, Whitlock J Jr. 1997. A novel cytoplasmic protein that interacts with the Ah receptor, con- tains tetratricopeptide repeat motifs, and augments the transcriptional response to 2,3,7,8- tetrachlorodibenzo-p-dioxin. Journal of Biological Chemistry 272(14):8878–8884. Marquez-Bravo LG, Gierthy JF. 2008. Differential expression of estrogen receptor alpha (ERalpha) protein in MCF-7 breast cancer cells chronically exposed to TCDD. Journal of Cellular Bio- chemistry 103(2):636–647. Martinez JM, Afshari CA, Bushel PR, Masuda A, Takahashi T, Walker NJ. 2002. Differential toxi- cogenomic responses to 2,3,7,8-tetrachlorodibenzo-p-dioxin in malignant and nonmalignant human airway epithelial cells. Toxicological Sciences 69(2):409–423. McCarty KM, Chen YC, Quamruzzaman Q, Rahman M, Mahiuddin G, Hsueh YM, Su L, Smith T, Ryan L, Christiani DC. 2007. Arsenic methylation, GSTT1, GSTM1, GSTP1 polymorphisms, and skin lesions. Environmental Health Perspectives 115(3):341–345. McMillan BJ, Bradfield CA. 2007. The aryl hydrocarbon receptor is activated by modified low- density lipoprotein. Proceedings of the National Academy of Sciences of the United States of America 104(4):1412–1417. Meza M, Gandolfi AJ, Klimecki WT. 2007. Developmental and genetic modulation of arsenic bio- transformation: A gene by environment interaction? Toxicology and Applied Pharmacology 222(3):381–387. Michalek JE, Pirkle JL, Needham LL, Patterson DG Jr, Caudill SP, Tripathi RC, Mocarelli P. 2002. Pharmacokinetics of 2,3,7,8-tetrachlorodibenzo-p-dioxin in Seveso adults and veterans of Operation Ranch Hand. Journal of Exposure Analysis and Environmental Epidemiology 12(1):44–53. Mitchell KA, Lockhart CA, Huang G, Elferink CJ. 2006. Sustained aryl hydrocarbon receptor activity attenuates liver regeneration. Molecular Pharmacology 70(1):163–170. Mizoi M, Takabayashi F, Nakano M, An Y, Sagesaka Y, Kato K, Okada S, Yamanaka K. 2005. The role of trivalent dimethylated arsenic in dimethylarsinic acid-promoted skin and lung tumori- genesis in mice: Tumor-promoting action through the induction of oxidative stress. Toxicology Letters 158(2):87–94. Mouron SA, Grillo CA, Dulout FN, Golijow CD. 2005. DNA-protein cross-links and sister chroma- tid exchanges induced by dimethylarsinic acid in human fibroblasts cells. Mutation Research 581(1-2):83–90. Naranmandura H, Suzuki KT. 2008. Formation of dimethylthioarsenicals in red blood cells. Toxicol- ogy and Applied Pharmacology 227(3):390–399. NCI (National Cancer Institute). 1978. Bioassay of Picloram for Possible Carcinogenicity. Report Series No. 23. Bethesda, MD: NCI. Nebert DW, Negishi M, Lang M, Hjelmeland L, Eisen H. 1982. The Ah locus, a multigene family necessary for survival in a chemically adverse environment: Comparison with the immune system. Advances in Genetics 21:1–52. Nebert DW, Roe AL, Dieter MZ, Solis WA, Yang Y, Dalton TP. 2000. Role of the aromatic hydro- carbon receptor and [Ah] gene battery in the oxidative stress response, cell cycle control, and apoptosis. Biochemical Pharmacology 59(1):65–85.

100 VETERANS AND AGENT ORANGE: UPDATE 2008 Needham L, Gerthoux P, Patterson D, Brambilla P, Pirkle J, Tramacere P, Turner W, Beretta C, Sampson E, Mocarelli P. 1994. Half-life of 2,3,7,8-tetrachlorodibenzo-p-dioxin in serum of Seveso adults: Interim report. Organohalogen Compounds 21:81–85. Neubert D, Wiesmuller T, Abraham K, Krowke R, Hagenmaier H. 1990. Persistence of various poly- chlorinated dibenzo-p-dioxins and dibenzofurans (PCDDs and PCDFs) in hepatic and adipose tissue of marmoset monkeys. Archives of Toxicology 64:431–442. Nolan R, Smith F, Hefner J. 1979. Elimination and tissue distribution of 2,3,7,8-tetrachlorodibenzo- para-dixion (TCDD) in female guinea-pigs following a single oral dose. Toxicology and Applied Pharmacology 48(1):A162. Nolan RJ, Freshour NL, Kastl PE, Saunders JH. 1984. Pharmacokinetics of picloram in male volun- teers. Toxicology and Applied Pharmacology 76:264–269. NRC (National Research Council). 1999a. Arsenic in Drinking Water. Washington, DC: National Academy Press. NRC. 2001. Arsenic in Drinking Water: Update 2001. Washington, DC: National Academy Press. Oesch-Bartlomowicz B, Huelster A, Wiss O, Antoniou-Lipfert P, Dietrich C, Arand M, Weiss C, Bockamp E, Oesch F. 2005. Aryl hydrocarbon receptor activation by cAMP vs. dioxin: Diver- gent signaling pathways. Proceedings of the National Academy Sciences of the United States of America 102:9218–9223. Pegram RA, Diliberto JJ, Moore TC, Gao P, Birnbaum LS. 1995. 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) distribution and cytochrome P4501A induction in young adult and senescent male mice. Toxicology Letters 76:119–126. Petrulis J, Kusnadi A, Ramadoss P, Hollingshead B, Perdew G. 2003. The hsp90 co-chaperone XAP2 alters importin beta recognition of the bipartite nuclear localization signal of the Ah receptor and represses transcriptional activity. Journal of Biological Chemistry 278:2677–2685. Piaggi S, Novelli M, Martino L, Masini M, Raggi C, Orciuolo E, Masiello P, Casini A, De Tata V. 2007. Cell death and impairment of glucose-stimulated insulin secretion induced by 2,3,7,8- tetrachlorodibenzo-p-dioxin (TCDD) in the beta-cell line INS-1E. Toxicology and Applied Pharmacology 220(3):333–340. Pirkle J, Wolfe W, Patterson D, Needham L, Michalek J, Miner J, Peterson M, Phillips D. 1989. Es- timates of the half-life of 2,3,7,8-tetrachlorodibenzo-p-dioxin in Vietnam veterans of Operation Ranch Hand. Journal of Toxicology and Environmental Health 27:165–171. Pohjanvirta R, Tuomisto J. 1990. Remarkable residual alterations in responses to feeding regulatory challenges in Han/Wistar rats after recovery from the acute toxicity of 2,3,7,8-tetrachlorodibenzo- p-dioxin (TCDD). Food and Chemical Toxicology 28:677–686. Poiger H, Schlatter C. 1986. Pharmacokinetics of 2,3,7,8-TCDD in man. Chemosphere 15: 1489–1494. Poland A, Glover E, Kende A. 1976. Stereospecific, high affinity binding of 2,3,7,8-tetrachlorodibenzo- p-dioxin by hepatic cytosol. Evidence that the binding species is receptor for induction of aryl hydrocarbon hydroxylase. Journal of Biological Chemistry 251:4936–4946. Poland A, Teitelbaum P, Glover E. 1989. [125I]2-iodo-3,7,8-trichlorodibenzo-p-dioxin-binding spe- cies in mouse liver induced by agonists for the Ah receptor: Characterization and identification. Molecular Pharmacology 36(1):113–120. Pongratz I, Mason GGF, Poellinger L. 1992. Dual roles of the 90-kDa heat shock protein hsp90 in modulating functional activities of the dioxin receptor. Journal of Biological Chemistry 267:13728–13734. Pont AR, Charron AR, Brand RM. 2004. Active ingredients in sunscreens act as topical penetration enhancers for the herbicide 2,4-dichlorophenoxyacetic acid. Toxicology and Applied Pharma- cology 195:348–354. Probst MR, Reisz-Porszasz S, Agbunag RV, Ong MS, Hankinson O. 1993. Role of the aryl hy- drocarbon receptor nuclear translocator protein in aryl hydrocarbon (dioxin) receptor action. Molecular Pharmacology 44:511–518.

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 101 Puga A, Bohm J, Hoffer A, Leikauf G, Shertzer H, Zhou S. 1995. Dioxin alters calcium homeostasis and the regulation of arachidonate metabolism in mouse hepatoma cells. Proceeding of the 15th International Symposium on Chlorinated Dioxins and Related Compounds 25:381–386. Puga A, Maier A, Medvedovic M. 2000. The transcriptional signature of dioxin in human hepatoma hepg2 cells. Biochemical Pharmacology 60(8):1129–1142. Puga A, Sartor MA, Huang MY, Kerzee JK, Wei YD, Tomlinson CR, Baxter CS, Medvedovic M. 2004. Gene expression profiles of mouse aorta and cultured vascular smooth muscle cells differ widely, yet show common responses to dioxin exposure. Cardiovascular Toxicology 4(4):385–404. Qin H, Powell-Coffman JA. 2004. The Caenorhabditis elegans aryl hydrocarbon receptor, AHR-1, regulates neuronal development. Developmental Biology 270(1):64–75. Quintana FJ, Basso AS, Iglesias AH, Korn T, Farez MF, Bettelli E, Caccamo M, Oukka M, Weiner HL. 2008. Control of T(reg) and T(H)17 cell differentiation by the aryl hydrocarbon receptor. Nature 453(7191):65–71. Rizki M, Kossatz E, Velazquez A, Creus A, Farina M, Fortaner S, Sabbioni E, Marcos R. 2006. Me- tabolism of arsenic in Drosophila melanogaster and the genotoxicity of dimethylarsinic acid in the Drosophila wing spot test. Environmental and Molecular Mutagenesis 47(3):162–168. Rose JQ, Ransy JC, Wentzler TH, Hummel RA, Gehring PJ. 1976. The fate of 2,3,7,8-tetrachloro- dibenzo-p-dioxin following single and repeated oral doses to the rat. Toxicology and Applied Pharmacology 36:209–226. Sato S, Shirakawa H, Tomita S, Ohsaki Y, Haketa K, Tooi O, Santo N, Tohkin M, Furukawa Y, Gonzalez FJ, Komai M. 2008. Low-dose dioxins alter gene expression related to cholesterol bio- synthesis, lipogenesis, and glucose metabolism through the aryl hydrocarbon receptor-mediated pathway in mouse liver. Toxicology and Applied Pharmacology 229(1):10–19. Sauerhoff MW, Braun WH, Blau GE, Gehring PJ. 1977. The fate of 2,4-dichlorophenoxyacetic acid (2,4-D) following oral administration to man. Toxicology 8:3–11. Savouret J, Antenos M, Quesne M, Xu J, Milgrom E, Casper R. 2001. 7-ketocholesterol is an en- dogenous modulator for the arylhydrocarbon receptor. Journal of Biological Chemistry (276): 3054–3059. Schmidt JV, Su GH, Reddy JK, Simon MC, Bradfield CA. 1996. Characterization of a murine Ahr null allele: Involvement of the Ah receptor in hepatic growth and development. Proceedings of the National Academy of Sciences of the United States of America 93(13):6731–6736. Schnekenburger M, Peng L, Puga A. 2007. HDAC1 bound to the Cyp1a1 promoter blocks histone acetylation associated with Ah receptor-mediated trans-activation. Biochimica et Biophysica ACTA/General Subjects 1769:569–578. Seidel S, Winters G, Rogers W, Ziccardi M, Li V, Keser B, Denison MS. 2001. Activation of the Ah receptor signaling pathway by prostaglandins. Journal of Biochemistry and Molecular Toxicol- ogy 15:187–196. Sen B, Wang A, Hester SD, Robertson JL, Wolf DC. 2005. Gene expression profiling of responses to dimethylarsinic acid in female F344 rat urothelium. Toxicology 215(3):214–226. Senft AP, Dalton TP, Nebert DW, Genter MB, Puga A, Hutchinson RJ, Kerzee JK, Uno S, Shertzer HG. 2002. Mitochondrial reactive oxygen production is dependent on the aromatic hydrocarbon receptor. Free Radical Biology and Medicine 33(9):1268–1278. Shen ES, Whitlock JP. 1992. Protein–DNA interactions at a dioxin responsive enhancer: Mutational analysis of the DNA binding site for the liganded Ah receptor. Journal of Biological Chemistry 267:6815–6819. Shertzer HG, Genter MB, Shen D, Nebert DW, Chen Y, Dalton TP. 2006. TCDD decreases ATP lev- els and increases reactive oxygen production through changes in mitochondrial F(0)F(1)-ATP synthase and ubiquinone. Toxicology and Applied Pharmacology 217(3):363–374. Sinal C, Bend J. 1997. Aryl hydrocarbon receptor–dependent induction of cyp1a1 by bilirubin in mouse hepatoma hepa 1c1c7 cells. Molecular Pharmacology 52:590–599.

102 VETERANS AND AGENT ORANGE: UPDATE 2008 Singh NP, Nagarkatti M, Nagarkatti P. 2008. Primary peripheral T cells become susceptible to 2,3,7,8- tetrachlorodibenzo-p-dioxin–mediated apoptosis in vitro upon activation and in the presence of dendritic cells. Molecular Pharmacology 73(6):1722–1735. Sogawa K, Numayama-Tsuruta K, Takahashi T, Matsushita N, Miura C, Nikawa J, Gotoh O, Kikuchi Y, Fujii-Kuriyama Y. 2004. A novel induction mechanism of the rat CYP1A2 gene mediated by Ah receptor–Arnt heterodimer. Biochemical and Biophysical Research Communications 318(3):746–755. Song Z, Pollenz RS. 2002. Ligand-dependent and independent modulation of aryl hydrocarbon recep- tor localization, degradation, and gene regulation. Molecular Pharmacology 62(4):806–816. Sonneveld E, Jonas A, Meijer OC, Brouwer A, van der Burg B. 2007. Glucocorticoid-enhanced expression of dioxin target genes through regulation of the rat aryl hydrocarbon receptor. Toxi- cological Sciences 99(2):455–469. Steinmaus C, Moore LE, Shipp M, Kalman D, Rey OA, Biggs ML, Hopenhayn C, Bates MN, Zheng S, Wiencke JK, Smith AH. 2007. Genetic polymorphisms in MTHFR 677 and 1298, GSTM1 and T1, and metabolism of arsenic. Journal of Toxicology and Environmental Health, Part A 70(2):159–170. Stott WT, Johnson KA, Landry TD, Gorzinski SJ, Cieszlak FS. 1990. Chronic toxicity and onco- genicity of picloram in Fischer 344 rats. Journal of Toxicology and Environmental Health 30:91–104. Suzuki KT, Katagiri A, Sakuma Y, Ogra Y, Ohmichi M. 2004. Distributions and chemical forms of arsenic after intravenous administration of dimethylarsinic and monomethylarsonic acids to rats. Toxicology and Applied Pharmacology 198(3):336–344. Swanson HI, Chan WK, Bradfield CA. 1995. DNA binding specificities and pairing rules of the Ah receptor, ARNT, and SIM proteins. Journal of Biological Chemistry 270(44):26292–26302. Tanaka J, Yonemoto J, Zaha H, Kiyama R, Sone H. 2007. Estrogen-responsive genes newly found to be modified by TCDD exposure in human cell lines and mouse systems. Molecular and Cellular Endocrinology 272(1-2):38–49. Thompson DJ, Emerson JL, Strebing RJ, Gerbig CG, Robinson VB. 1972. Teratology and postnatal studies on 4-amino-3,5,6-trichloropicolinic acid (picloram) in the rat. Food and Cosmetics Toxicology 10:797–803. Timchalk C. 2004. Comparative inter-species pharmacokinetics of phenoxyacetic acid herbicides and related organic acids. Evidence that the dog is not a relevant species for evaluation of human health risk. Toxicology 200(1):1–19. Tonack S, Kind K, Thompson JG, Wobus AM, Fischer B, Santos AN. 2007. Dioxin affects glucose transport via the arylhydrocarbon receptor signal cascade in pluripotent embryonic carcinoma cells. Endocrinology 148(12):5902–5912. Tsuchiya M, Katoh T, Motoyama H, Sasaki H, Tsugane S, Ikenoue T. 2005. Analysis of the AhR, ARNT, and AhRR gene polymorphisms: Genetic contribution to endometriosis susceptibility and severity. Fertility and Sterility 84(2):454–458. Vahter M, Concha G. 2001. Role of metabolism in arsenic toxicity. Pharmacology and Toxicology 89(1):1–5. Vahter M, Marafante E, Dencker L. 1984. Tissue distribution and retention of 74As-dimethylarsinic acid in mice and rats. Archives of Environmental Contamination and Toxicology 13:259–264. Veldhoen M, Hirota K, Westendorf AM, Buer J, Dumoutier L, Renauld JC, Stockinger B. 2008. The aryl hydrocarbon receptor links TH17-cell-mediated autoimmunity to environmental toxins. Nature 453(7191):106–109. Vezina CM, Walker NJ, Olson JR. 2004. Subchronic exposure to TCDD, PeCDF, PCB126, and PCB153: Effect on hepatic gene expression. Environmental Health Perspectives 112(16): 1636–1644. Viluksela M, Duong TV, Stahl BU, Li X, Tuomisto J, Rozman KK. 1996. Toxicokinetics of 2,3,7,8- tetrachlorodibenzo-p-dioxin (TCDD) in two substrains of male Long-Evans rats after intrave- nous injection. Fundamental and Applied Toxicology 31:184–191.

INFORMATION RELATED TO BIOLOGIC PLAUSIBILITY 103 Vogel CF, Sciullo E, Li W, Wong P, Lazennec G, Matsumura F. 2007a. RelB, a new partner of aryl hydrocarbon receptor–mediated transcription. Molecular Endocrinology 21(12):2941–2955. Vogel CF, Sciullo E, Matsumura F. 2007b. Involvement of RelB in aryl hydrocarbon receptor- mediated induction of chemokines. Biochemical and Biophysical Research Communications 363(3):722–726. Vogel CF, Goth SR, Dong B, Pessah IN, Matsumura F. 2008. Aryl hydrocarbon receptor signaling mediates expression of indoleamine 2,3-dioxygenase. Biochemical and Biophysical Research Communications 375(3):331–335. Weber LW, Zesch A, Rozman K. 1991. Penetration, distribution and kinetics of 2,3,7,8-tetrachlorod- ibenzo-p-dioxin in human skin in vitro. Archives of Toxicology 65:421–428. Weber LW, Ernst SW, Stahl BU, Rozman K. 1993. Tissue distribution and toxicokinetics of 2,3,7,8- tetrachlorodibenzo-p-dioxin in rats after intravenous injection. Fundamental and Applied Toxi- cology 21:523–534. Wei M, Wanibuchi H, Morimura K, Iwai S, Yoshida K, Endo G, Nakae D, Fukushima S. 2002. Carci- nogenicity of dimethylarsinic acid in male F344 rats and genetic alterations in induced urinary bladder tumors. Carcinogenesis 23(8):1387–1397. Wei M, Arnold L, Cano M, Cohen SM. 2005. Effects of co-administration of antioxidants and arseni- cals on the rat urinary bladder epithelium. Toxicological Sciences 83(2):237–245. Weiss C, Faust D, Schreck I, Ruff A, Farwerck T, Melenberg A, Schneider S, Oesch-Bartlomowicz B, Zatloukalova J, Vondracek J, Oesch F, Dietrich C. 2008. TCDD deregulates contact inhibition in rat liver oval cells via Ah receptor, JunD and cyclin A. Oncogene 27(15):2198–2207. Whitelaw M, Pongratz I, Wilhelmsson A, Gustafsson JA, Poellinger L. 1993. Ligand-dependent recruitment of the ARNT coregulatory determines DNA recognition by the dioxin receptor. Molecular and Cellular Biology 13:2504–2514. WHO (World Health Organization). 1984. 2,4-Dichlorophenoxyacetic Acid (2,4-D). Environmental Health Criteria 29. Geneva: WHO. Xie Y, Trouba KJ, Liu J, Waalkes MP, Germolec DR. 2004. Biokinetics and subchronic toxic effects of oral arsenite, arsenate, monomethylarsonic acid, and dimethylarsinic acid in v-Ha-ras transgenic (Tg.AC) mice. Environmental Health Perspectives 112(12):1255–1263. Yao E, Denison M. 1992. DNA sequence determinants for binding of transformed Ah receptor to a dioxin-responsive enhancer. Biochemistry 31:5060–5067. Zarzour AH, Selim M, Abd-Elsayed AA, Hameed DA, AbdelAziz MA. 2008. Muscle invasive blad- der cancer in Upper Egypt: The shift in risk factors and tumor characteristics. BMC Cancer 8(250).

Next: 5 Epidemiologic Studies - New Citations and Background on Repeatedly Studied Populations »
Veterans and Agent Orange: Update 2008 Get This Book
×
Buy Hardback | $140.00 Buy Ebook | $109.99
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

From 1962 to 1971, the U.S. military sprayed herbicides over Vietnam to strip the thick jungle canopy that could conceal opposition forces, to destroy crops that those forces might depend on, and to clear tall grasses and bushes from the perimeters of U.S. base camps and outlying fire-support bases.

In response to concerns and continuing uncertainty about the long-term health effects of the sprayed herbicides on Vietnam veterans, Veterans and Agent Orange provides a comprehensive evaluation of scientific and medical information regarding the health effects of exposure to Agent Orange and other herbicides used in Vietnam. The 2008 report is the eighth volume in this series of biennial updates. It will be of interest to policy makers and physicians in the federal government, veterans and their families, veterans' organizations, researchers, and health professionals.

  1. ×

    Welcome to OpenBook!

    You're looking at OpenBook, NAP.edu's online reading room since 1999. Based on feedback from you, our users, we've made some improvements that make it easier than ever to read thousands of publications on our website.

    Do you want to take a quick tour of the OpenBook's features?

    No Thanks Take a Tour »
  2. ×

    Show this book's table of contents, where you can jump to any chapter by name.

    « Back Next »
  3. ×

    ...or use these buttons to go back to the previous chapter or skip to the next one.

    « Back Next »
  4. ×

    Jump up to the previous page or down to the next one. Also, you can type in a page number and press Enter to go directly to that page in the book.

    « Back Next »
  5. ×

    To search the entire text of this book, type in your search term here and press Enter.

    « Back Next »
  6. ×

    Share a link to this book page on your preferred social network or via email.

    « Back Next »
  7. ×

    View our suggested citation for this chapter.

    « Back Next »
  8. ×

    Ready to take your reading offline? Click here to buy this book in print or download it as a free PDF, if available.

    « Back Next »
Stay Connected!