The Risk Assessment Information System


  Toxicity
Values
Toxicity
Profiles
PRGs Risk
Models
Ecological
Benchmarks
ARARs Background
Values
Glossary

  RAIS Home

  What's New

  FAQ

  Tutorials

  Contacts

  Risk Assessment
   Guidance

  Regulatory
   Guidance

  Risk Tools

  ORO Site-Specific
   Information

  SADA

  EPA Tools

  Affiliated
   Organizations

 

 


Toxicity Profiles

Toxicity Summary for SELENIUM

NOTE: Although the toxicity values presented in these toxicity profiles were correct at the time they were produced, these values are subject to change. Users should always refer to the Toxicity Value Database for the current toxicity values.

Download a WordPerfect version of this toxicity profile. Please note that this document has been saved in WordPerfect 5.1/5.2 for greater accessibility but may have been originally formatted in later versions of WordPerfect (i.e., WordPerfect 6.1, Suite 7, etc.); therefore, formatting changes (i.e., Contents and Page Numbering) may occur when downloading this document.
EXECUTIVE SUMMARY
1. INTRODUCTION
2. METABOLISM AND DISPOSITION
2.1 ABSORPTION
2.2 DISTRIBUTION
2.3 METABOLISM
2.4 EXCRETION
3. NONCARCINOGENIC HEALTH EFFECTS
3.1 ORAL EXPOSURES
3.2 INHALATION EXPOSURES
3.3 OTHER ROUTES OF EXPOSURE
3.4 TARGET ORGANS/CRITICAL EFFECTS
4. CARCINOGENICITY
4.1 ORAL EXPOSURES
4.2 INHALATION EXPOSURES
4.3 OTHER ROUTES OF EXPOSURE
4.4 EPA WEIGHT-OF-EVIDENCE
4.5 CARCINOGENICITY SLOPE FACTORS
5. REFERENCES

MARCH 1993

Prepared by: Dennis M. Opresko, Ph.D, Chemical Hazard Evaluation Group, Biomedical Environmental Information Analysis Section, Health and Safety Research Division, Oak Ridge National Laboratory*, Oak Ridge, Tennessee.

Prepared for: Oak Ridge Reservation Environmental Restoration Program.

*Managed by Martin Marietta Energy Systems, Inc., for the U.S. Department of Energy under Contract No. DE-AC05-84OR21400.

EXECUTIVE SUMMARY

Selenium is an essential trace element important in many biochemical and physiological processes including the biosynthesis of coenzyme Q (a component of mitochondrial electron transport systems), regulation of ion fluxes across membranes, maintenance of the integrity of keratins, stimulation of antibody synthesis, and activation of glutathione peroxidase (an enzyme involved in preventing oxidative damage to cells). Recommended human dietary allowances (average daily intake) for selenium are as follows: infants up to 1 year, 10-15 µg; children 1-10 years, 20-30 µg; adult males 11-51+ years, 40-70 µg; adult females 11-51+ years, 45-55 µg; pregnant or lactating women, 65-75 µg. There appears to be a relatively narrow range between levels of selenium intake resulting in deficiency and those causing toxicity.

Selenium occurs in several valence states: -2 (hydrogen selenide, sodium selenide, dimethyl selenium, trimethyl selenium, and selenoamino acids such as selenomethionine; 0 (elemental selenium); +4 (selenium dioxide, selenious acid, and sodium selenite); and +6 (selenic acid and sodium selenate). Toxicity of selenium varies with valence state and water solubility of the compound in which it occurs. The latter can affect gastrointestinal absorption rates.

Gastrointestinal absorption in animals and humans for various selenium compounds ranges from about 44% to 95% of the ingested dose (Thomson and Stewart, 1974; Bopp et al., 1982; Thomson, 1974). Respiratory tract absorption rates of 97% and 94% for aerosols of selenious acid have been reported for dogs and rats, respectively (Weissman et al., 1983; Medinsky et al., 1981). Selenium is found in all tissues of the body; highest concentrations occur in the kidney, liver, spleen, and pancreas (Schroeder and Mitchener, 1971a; Schroeder and Mitchener, 1972; Jacobs and Forst, 1981a; Julius et al., 1983; Shamberger, 1984; Echevarria et al., 1988). Excretion is primarily via the urine (0-15 µg/L); however, excretory products can also be found in the feces, sweat, and in expired air.

In humans, acute oral exposures can result in excessive salivation, garlic odor to the breath, shallow breathing, diarrhea, pulmonary edema, and death (Civil and McDonald, 1978; Carter, 1966; Koppel et al., 1986). Other reported signs and symptoms of acute selenosis include tachycardia, nausea, vomiting, abdominal pain, abnormal liver function, muscle aches and pains, irritability, chills, and tremors. Acute toxic effects observed in animals include pulmonary congestion, hemorrhages and edema, convulsions, altered blood chemistry (increased hemoglobin and hematocrit); liver congestion; and congestion and hemorrhage of the kidneys (Smith et al., 1937; Anderson and Moxon, 1942; Hopper et al., 1985).

General signs and symptoms of chronic selenosis in humans include loss of hair and nails, acropachia (clubbing of the fingers), skin lesions (redness, swelling, blistering, and ulcerations), tooth decay (mottling, erosion and pitting), and nervous system abnormalities attributed to polyneuritis (peripheral anesthesia, acroparaethesia, pain in the extremities, hyperreflexia of the tendon, numbness, convulsions, paralysis, motor disturbances, and hemiplegia). In domesticated animals, subchronic and chronic oral exposures can result in loss of hair, malformed hooves, rough hair coat, and nervous system abnormalities (impaired vision and paralysis). Damage to the liver and kidneys and impaired immune responses have been reported to occur in rodents following subchronic and/or chronic oral exposures (Ganther and Baumann, 1962; Beems and van Beek, 1985; NCI, 1980a; Tinsley et al., 1967; Harr et al., 1967; Schroeder, 1967).

Selenium is teratogenic in birds and possibly also in domesticated animals (pigs, sheep, and cattle), but evidence of teratogenicity in humans and laboratory animals is lacking (ASTDR, 1989). However, adverse reproductive and developmental effects (decreased rates of conception, increased rates of fetal resorption, and reduced fetal body weights) have been reported for domesticated and laboratory animals (Harr and Muth, 1972: Wahlstrom and Olson, 1959; Schroeder and Mitchener, 1971b).

The Reference Dose (RfD) for chronic oral exposures is 0.005 mg/kg/day for both selenium and selenious acid (U.S. EPA, 1992a, 1992b). The subchronic RfDs for these compounds are the same as the chronic RfDs (U.S. EPA, 1992c).

In humans, inhalation of selenium or selenium compounds primarily affects the respiratory system. Dusts of elemental selenium and selenium dioxide can cause irritation of the skin and mucous membranes of the nose and throat, coughing, nosebleed, loss of sense of smell, dyspnea, bronchial spasms, bronchitis, and chemical pneumonia (Clinton, 1947; Hamilton, 1949). Other signs and symptoms following acute inhalation exposures include lacrimation, irritation and redness of the eyes, gastrointestinal distress (nausea and vomiting), depressed blood pressure, elevated pulse rate, headaches, dizziness, and malaise (ATSDR, 1989). In animals, acute inhalation exposures also result in severe respiratory effects including edema, hemorrhage, and interstitial pneumonitis (Hall et al., 1951; Dudley and Miller, 1937) as well as in splenic damage (congestion, fissuring red pulp, and increased polymorphonuclear leukocytes) and liver congestion and mild central atrophy (Hall et al., 1951). Information on toxicity of selenium in humans and animals following chronic inhalation exposures is not available, and subchronic and chronic inhalation Reference Concentrations have not been derived.

Epidemiologic studies in humans havation between chronic oral exposures to selenium and an increased incidence of death due to neoplasms. Some studies have indicated that selenium may have anti-neoplastic properties (see Whanger, 1983; Hocman, 1988). In studies on laboratory animals, selenites or selenates have not been found to be carcinogenic; however, selenium sulfide produced a significant increase in the incidence of hepatocellular carcinomas in male and female rats and in female mice and a significant increase in alveolar/bronchiolar carcinomas and adenomas in female mice following chronic oral exposures (NCI, 1980c). EPA has placed selenium and selenious acid in Group D, not classifiable as to carcinogenicity in humans (U.S. EPA, 1992a and 1992b), while selenium sulfide is placed in Group B2, probable human carcinogen (U.S. EPA, 1992d). Quantitative data are, however, insufficient to derive a slope factor for selenium sulfide. Pertinent data regarding the potential carcinogenicity of selenium by the inhalation route in humans or animals were not located in the available literature.

1. INTRODUCTION

Selenium is an essential trace element important in many biochemical and physiological processes including the biosynthesis of coenzyme Q (a component of mitochondrial electron transport systems), regulation of ion fluxes across membranes, maintenance of the integrity of keratins, stimulation of antibody synthesis, and activation of glutathione peroxidase (an enzyme involved in preventing oxidative damage to cells) (Hammond and Beliles, 1980). Animal studies indicate that deficiencies in selenium can result in damage to the liver, heart, kidneys, skeletal muscle, and testes (Hammond and Beliles, 1980). Recommended human dietary allowances (average daily intake) for selenium are as follows: infants up to 1 year, 10-15 µg; children 1-10 years, 20-30 µg; adult males 11-51+ years, 40-70 µg; adult females 11-51+ years, 45-55 µg; pregnant or lactating women, 65-75 µg. The primary dietary sources of selenium are seafoods, kidney and liver meats, and grains and cereals (NRC, 1989).

Selenium occurs in several oxidation states: -2 (hydrogen selenide, sodium selenide, dimethyl selenium, trimethyl selenium, and selenoamino acids such as selenomethionine; 0 (elemental selenium); +4 (selenium dioxide, selenious acid, and sodium selenite); and +6 (selenic acid and sodium selenate).

2. METABOLISM AND DISPOSITION

2.1. ABSORPTION

Gastrointestinal absorption in humans for various selenium compounds ranges from about 44% to 95% of the ingested dose (Thomson and Stewart, 1974; Bopp et al., 1982; Thomson, 1974). Absorption is highest when the compound is administered in solution and lowest when it is administered as a solid. Absorption is also more efficient after a single dose than after repetitive daily doses.

In studies on rats, mice and dogs, gastrointestinal absorption rates of 87% or more have been reported for [75Se]-selenite (selenious acid) (Bopp et al., 1982; U.S. EPA, 1989: Furchner et al.,1975). Absorption is highest following gavage administration, but may be only 50% when the compound is administered in feed (Weissmann et al., 1983).

Respiratory tract absorption rates of 97% and 94% for aerosols of [75Se]-selenite (selenious acid) have been reported for dogs and rats, respectively (Weissman et al., 1983; Medinsky et al., 1981).

2.2. DISTRIBUTION

Selenium is found in all tissues at concentrations that vary with the amount ingested in the diet and the type of tissue. Highest concentrations occur in the kidney, liver, spleen, and pancreas (Schroeder and Mitchener, 1971a; Schroeder and Mitchener, 1972; Jacobs and Forst, 1981a; Julius et al., 1983; Shamberger, 1984; Echevarria et al., 1988). Selenium is also concentrated in erythrocytes relative to the amount in blood plasma (Butler et al., 1990). As a result of occupational exposures, high concentrations can also be found in peribronchial nodes, lung, hair, and nails (Diskin et al., 1979).

2.3. METABOLISM

Two major metabolic products of selenite have been identified, dimethyl selenide and trimethylselenonium ion (Palmer et al., 1969, 1970; Nakamuro et al., 1977; Jiang et al., 1983). In the formation of dimethyl selenide, selenite (H2SeO3) is first reduced nonenzymatically to the stable selenotrisulfide (GSSeSG) by four glutathione molecules. An NADPH-dependent reduction involving glutathione reductase converts GSSeSG to the very unstable selenopersulfide (GSSeH), and a further reduction by NADPH and glutathione reductase converts GSSeH to hydrogen selenide (H2Se). Methyl groups donated by S-adenosylmethionine are transferred by a methyltransferase to hydrogen selenide to form dimethyl selenide (Ganther, 1979; see also Whiting et al., 1980 and Bopp et al., 1982). The reactions involved in the metabolism of selenite to trimethylselenonium are not known; however, according to Bopp et al. (1982), it is not a simple transfer of an additional methyl group to dimethyl selenide.

2.4. EXCRETION

Within certain physiological limits trace amounts of selenium are retained in the body while excessive amounts are excreted (Hammond and Beliles, 1980). Excretion is primarily in the urine; however, excretory products can also be found in the feces, sweat, and in expired air. At high dietary levels a significant fraction of the dose is eliminated in expired air, but at low dietary levels the amount in expired air is insignificant. Only a small fraction of the amount absorbed through the gastrointestinal tract is excreted in the feces; however, for selenium compounds that are not readily absorbed 33-58% of the dose may appear in the feces (Bopp et al., 1982; Shamberger, 1984; Thomson and Stewart, 1974).

Urinary excretion of selenium in humans varies with dose and frequency of administration. According to Bopp et al. (1982), urinary excretion after ingestion of trace amounts seldom exceeds 10-15%. Thomson and Stewart (1974) reported that only 10% of a 10 µg ingested dose of [75Se]-selenite was excreted in urine within 14 days; however, urinary excretion was 41-54% following a single 0.1-mg dose, 72-77% after a single 1-mg dose, and 36% after five 1-mg daily doses (by day 9) (Thomson et al., 1978). The elimination curve for selenium in humans is triphasic and the terminal half-life for selenite is about 4 months (Bopp et al., 1982).

Excretion of selenium in expired air of occupationally exposed individuals is very low to undetectable. Among individuals exposed to high acute levels a garlic-like odor to the breath is indicative of the presence of dimethylselenide.

Excretion of selenium in animals follows a pattern similar to that in humans. In tests on rats maintained on a diet supplemented with 1 ppm selenium, up to 67% was excreted in the urine but only 10% in the feces (Burk et al., 1972). Mean half-times of excretion ranged from 19.5 days for animals on a 0.1 ppm selenium diet to 1.2 days for animals on a 1 ppm diet. Elimination of selenium by way of the respiratory tract was usually <10% (Burk et al., 1972).

3. NONCARCINOGENIC HEALTH EFFECTS

3.1. ORAL EXPOSURES

3.1.1. Acute Toxicity

3.1.1.1. Human

Several cases of selenium poisoning in humans have been reported. A 3-yr-old boy died 1.5 hours after ingesting an undetermined amount of selenious acid (Carter 1966), and a 15-yr-old girl survived after ingesting an estimated 22.3 mg Se/kg as sodium selenate (Civil and MacDonald, 1978). Clinical signs included excessive salivation, garlic odor to the breath, shallow breathing, and diarrhea. Pulmonary edema and lung lesions have been reported in cases of acute lethal exposures (Carter, 1966; Koppel et al., 1986). Other reported signs and symptoms of acute selenosis include tachycardia, nausea, vomiting, abdominal pain, abnormal liver function, muscle aches and pains, irritability, chills, and tremors.

3.1.1.2. Animal

The acute oral toxicity of selenium varies with the solubility of the chemical compound in which it occurs; the more soluble compounds such as sodium selenite and sodium selenate are more toxic than the less soluble elemental selenium, selenium sulfide and selenium disulfide (ATSDR, 1989). Oral LD50 values for sodium selenite range from 1 to 7 mg Se/kg (rats, rabbits, mice, and guinea pigs), whereas an LD50 of 138 mg/kg has been reported for selenium disulfide, and a 10-d LD50 of 6,700 mg Se/kg has been reported for elemental selenium administered to rats (Cummins and Kimura, 1971; Pletnikova, 1970).

Respiratory effects in laboratory animals following acute oral exposures are similar to those occurring in humans and include pulmonary congestion, hemorrhages and edema, labored respiration, muscular weakness, and asphyxial convulsions (Smith et al., 1937). Acute exposures can also result in altered blood chemistry (increased hemoglobin and hematocrit in dogs); liver congestion; and congestion and hemorrhage of the kidneys (Anderson and Moxon, 1942; Hopper et al., 1985).

3.1.2. Subchronic Toxicity

3.1.2.1. Human

A 62-year old man ingesting 2 mg of sodium selenite daily (0.9 mg Se/day) for 2 years developed thickened and fragile nails and had a garlic-like odor to his perspiration (Yang et al., 1983). A 57-year old woman who had ingested an estimated 2,387 mg selenium over 77 days exhibited almost total alopecia, brittleness and loss of fingernails, sour-milk breath, nausea, vomiting and fatigue (CDC, 1984).

Listlessness, lack of mental alertness, and other symptoms of selenosis were observed in members of a family exposed to high levels of selenium in well water for about 3 months (Rosenfeld and Beath, 1964). The well water contained 9 mg Se/L and daily intake from drinking water was estimated to be 0.26 mg Se/kg/day. Dietary intake of selenium was not reported.

3.1.2.2. Animal

Livestock grazing on plants having a high selenium content develop a condition called "alkali disease", which is characterized by loss of hair, inflammation of the coronary band followed by cracked or malformed hooves, rough hair coat, impaired vision, aimless wandering behavior, reduced consumption of food and water, and paralysis (ATSDR, 1989). Symmetrical focal poliomyelomalacia has been observed in pigs exposed to selenium (Goehring et al., 1984; Harrison et al., 1983; Wilson and Drake, 1982), and in laboratory studies this syndrome was reproduced in pigs receiving 50 mg Se per kg of food for 20-40 days (Wilson et al., 1983). The estimated dose was 1.5 mg Se/kg body weight (ATSDR, 1989). Doses of 0.6 and 1.1 mg/kg/day to steers resulted in trembling and neural degeneration of the cerebral and cerebellar cortices (Maag et al., 1960).

A 41-48% depression in body weight gain (during the initial 4 weeks of treatment) was observed in male Holtzman rats maintained for 42-180 days on a diet containing 5 ppm Se (0.25 mg Se/kg/day) (Ganther and Baumann, 1962). By week 6, the relative weight of the spleen was increased twofold. Gross liver lesions, consisting of hobnailed or pebbled appearance, severe cirrhosis and necrosis, were also observed in the treated animals.

When administered to male Syrian hamsters for 42 days, sodium selenite, at a dietary level of 18 ppm Se (1.21 mg Se/kg/day), caused an 18% reduction in body weight gain and a 13% reduction in food intake (Beems and van Beek, 1985). Although no treatment-related gross lesions were seen, histopathological evaluation revealed oval cell proliferation in the periportal areas of the liver, enlargement of hepatocytes and their nuclei in the centrilobular area, focal necrosis, necrosis of individual cells, and pigment accumulation (lipofuscin and hemosiderin) in periportal macrophages (Beems and van Beek, 1985). No adverse effects occurred in animals receiving selenium at a dietary level of 9 ppm (0.61 mg Se/kg bw/day).

Focal coagulation necrosis occurred in the liver of rats dosed with 31.6 mg selenium sulfide/kg/day by gavage for 13 weeks, but not in rats receiving 17.8 mg selenium sulfide/kg/day (NCI, 1980a). A mixture of selenium sulfide and selenium disulfide, equivalent to a dose of 464 mg/kg/day, did not cause hepatic effects in mice treated by gavage for 13 weeks; however, the incidence and severity of interstitial nephritis increased compared to controls (NCI, 1980a). Renal effects did not occur in mice dosed with 216 mg/kg/day.

Reduced humoral antibody (IgC) production and reduced prostaglandin synthesis were observed in rats dosed for 10 weeks with 0.75 mg Se/kg/day (as sodium selenite in drinking water) (Koller et al., 1986). Lower doses (0.075 and 0.28 mg/kg/day) increased natural killer cell (NKC) cytotoxicity, enhancing the immune response to antigen stimulation; however, delayed-type hypersensitivity (DTH) and prostaglandin E2 activity were significantly reduced.

Ingestion of organic seleniferous wheat by young rats for six weeks resulted in a reduction in hemoglobin and the death of 1 of 8 animals receiving 0.68 mg Se/kg/day (Halverson et al., 1966). Growth was reduced 20% at 0.55 mg Se/kg/day, and the maximum reported NOAEL was 0.42 mg Se/kg/day.

3.1.3. Chronic Toxicity

3.1.3.1. Human

Selenium has been identified as the cause of the adverse health effects exhibited by the inhabitants of several villages in China in a region where coal with a high selenium content is used for cooking and also is burned in fields as a source of plant fertilizer (Yang et al., 1983). General signs and symptoms of selenosis included loss of hair and nails, acropachia (clubbing of the fingers), skin lesions (redness, swelling, blistering and ulcerations), tooth decay (mottling, erosion, and pitting), and nervous system abnormalities (peripheral anesthesia, acroparaethesia, pain in the extremities, hyperreflexia of the tendon, numbness, convulsions, paralysis, motor disturbances, and hemiplegia). The nervous system abnormalities were attributed to polyneuritis caused by exposure to selenium (Yang et al., 1983).

The daily selenium intake in the exposed population was estimated to be 4.99 mg (range = 3.2-6.69 mg/day). Selenium levels in hair, blood, and urine were 32.2 µg/g, 3.2 µg/mL and 2.68 µg/mL, respectively. For residents in a control area, selenium intake was 0.116 mg/day, and average concentrations in hair, blood, and urine were 0.36 µg/g, 0.095 µg/mL, and 0.026 µg/mL, respectively. Whanger (1989) suggested that selenosis in the endemic area may have been due, in part, to food contamination during cooking with seleniferous coal or from inhalation of volatilized selenium. Yang et al. (1983), however, did not consider inhalation of selenium as a contributing factor. The drinking water in the endemic area contained 54 µg Se/L.

Yang et al. (1989a,b) reevaluated the dietary intake of selenium using a larger group of subjects from low, medium, and high selenium areas. The daily selenium intake (mean±SE) for adults was as follows: males, 70.5±4.8 (low), 194.7±22.9 (medium), and 1438.2±76.4 µg/day (high); females, 62.0±3.6 (low), 198.1±23.8 (medium), 1238.5±64.6 µg/day (high). Selenium concentration in drinking water, 0.37, 1.72, and 12.27 µg/L for low, medium, and high areas, respectively, represented 3% or less of the daily intake (based on consumption of 3 L of water per day).

Blood selenium levels were obtained for five subjects presenting distinct and persisting clinical signs of selenosis; the levels ranged from 1.054-1.854 mg/L of blood, with a mean of 1.35 mg/L. Based on regression analysis [i.e., logSeblood = 0.767 logSeintake 2.248], the daily selenium intake corresponding to a blood level of 1.35 mg/L was 1,261 µg/day. Toxic signs of selenosis occurred most often in subjects with a previous history of selenosis. About 97% of the cases were adults older than 18 years and no cases occurred in children younger than 12 years, although children often had daily intakes and blood levels exceeding those of the adults. Clinical chemistry studies indicated a 44% increase in prothrombin time and a 19% reduction in whole blood glutathione at blood levels above 1 mg/L (850 µg/day). The white blood cell counts were also elevated in subjects from the high selenium area.

Longnecker et al. (1991) studied the selenium intake and health of 142 subjects living in western South Dakota and eastern Wyoming, a region where the soil has a high selenium content. Mean selenium intake was estimated to be 3.04 µmol/day (240 µg/day) with a range of 0.86-9.20 µmol/day (68-730 µg/day). No adverse health effects that might be associated with exposure to selenium were found in the study population.

3.1.3.2. Animal

The chronic oral toxicity of sodium selenite has been evaluated in several laboratory species. In one study, male and female Wistar rats were given dietary sodium selenite at concentrations of 0, 0.5, and 2.0 ppm in a semipurified high-protein diet for an entire lifetime (Tinsley et al., 1967; Harr et al., 1967). The maximum body weight was reduced by 9-10% in male and female rats receiving the 0.5-ppm diet and by 28% in rats receiving the 2-ppm diet. Histopathological effects consisted of hepatic parenchymal lesions and interstitial nephritis; both conditions were less severe and occurred less frequently in animals given sodium selenite than in those given sodium selenate. Myocarditis was more severe and occurred more frequently in animals given sodium selenite.

In a study conducted by Schroeder (1967), female Long-Evans strain rats were administered sodium selenite in drinking water for 24 months. The time-weighted-average (TWA) dose for the entire period was 0.173 mg Se/kg/day (0.379 mg sodium selenite/kg/day). Statistically significant reductions were observed in mean body weight (14% at 24 mo), mean heart weight at death, and survival rates (51%) of animals 9 months of age. Gross lesions observed in treated animals dying before 9 months of age were atrophy and irregular areas of fatty degeneration in the liver. Among those rats surviving until termination, one rat had focal cirrhosis of the liver. No other organs were affected (Schroeder, 1967; Schroeder and Mitchener, 1971a).

Schroeder (1967) also conducted a drinking water study using male and female Charles River CD mice exposed to sodium selenite from 20-22 days of age until 15 months of age. Concentrations in drinking water were 2 µg Se/mL [the weight-normalized dose was estimated to be 0.38 mg Se/kg/day (0.83 mg sodium selenite/kg/day)]. Gross lesions (not otherwise specified) were seen in the liver of 15 of 35 mice (sex not specified).

In another study, Schroeder and Mitchener (1972) administered 3 ppm of selenium (sodium selenite) in drinking water to male and female Charles River CD mice for their entire lifetime. Based on water consumption of 19% of body weight (U.S. EPA, 1986), the weight-normalized dose was 0.57 mg Se/kg/day (1.25 mg sodium selenite/kg/day). Treated mice were less active, in poor health, and shortly before death their extremities became edematous. Growth and longevity were enhanced in treated males but depressed in females. Amyloidosis of the kidney, heart, adrenal, liver, spleen, and lung was seen in 58% of the selenium-treated mice, but also in 30% of the controls.

Jacobs and Forst (1981a) administered 0 or 4 ppm of selenium (sodium selenite) to male Sprague-Dawley rats in their drinking water for 1 or 2 years. The estimated weight-normalized dose was 0.56 mg Se/kg/day (1.23 mg sodium selenite/kg/day). No treatment-related gross or histopathologic effects occurred in either treated group. Hepatic glutathione peroxidase activity was reduced by 50% in rats given selenium for only 1 year. In a related study, Jacobs and Forst (1981b) exposed female Swiss albino mice to selenium as sodium selenite (0, 1, 4, 8 ppm) in drinking water for 50 weeks. Based on data provided by the authors, weight-normalized doses were 0, 0.27, 0.84 and 1.41 mg Se/kg/day (0, 0.59, 1.83 and 3.09 mg sodium selenite/kg/day). The only reported treatment-related effect was a 50% reduction in weight gain in the mice receiving 8 ppm. Hepatic glutathione peroxidase activity was increased.

3.1.4. Developmental and Reproductive Toxicity

3.1.4.1. Human

There is no clinical or epidemiologic evidence that selenium is teratogenic in humans, and in one epidemiologic study of a population in which the incidence of selenosis was high, no selenium-associated reproductive or developmental effects were reported (Yang et al., 1989b).

3.1.4.2. Animal

Selenium is teratogenic in birds and possibly also in domesticated animals (pigs, sheep, and cattle), but evidence of teratogenicity in laboratory animals is lacking (ASTDR, 1989). Adverse developmental and reproductive effects have been reported in both domesticated and laboratory animals.

Decreased rates of conception and increased rates of fetal resorption have been reported in cattle, sheep, and horses fed diets high in organic selenium (25-50 mg/kg diet) (Harr and Muth, 1972). Similar effects on conception rates have been observed in pigs maintained on a diet containing the equivalent of 0.41 mg Se/kg/day (as sodium selenite) (Wahlstrom and Olson, 1959).

In 3-generation studies conducted on mice by Schroeder and Mitchener (1971b), sodium selenate (0.42 mg Se/kg/day in drinking water) had no adverse effects on reproduction, while sodium selenite (0.42 mg Se/kg/day) adversely affected breeding in 50% of the treated pairs.

Nobunaga et al. (1979) studied the maternal and fetal toxicity of sodium selenite administered in drinking water to female IVCS mice. Fourteen females received 11.4 nmol/mL (3 ppm; low dose) and ten females received 22.8 nmol/mL (6 ppm; high dose) 30 days before mating and through day 18 of gestation. The length of the estrus cycle was changed in about 12% of mice receiving the low dose. No significant effects were observed on the mean litter size or the numbers of implants, resorptions, dead fetuses or gross malformations. The mean fetal weight was reduced by 7% (p<0.01) at the high dose.

Hardin et al. (1987) dosed female CD-1 mice by gavage with sodium selenite dissolved in water. Fifty females per group were given 0, 3.5, 5, 7 or 14 mg sodium selenite/kg/day on gestation days 6-13 inclusive. No toxicity was observed in either dams or pups in the 3.5- to 7-mg/kg dose groups. The 14-mg/kg dose caused 44% maternal mortality (compared with 0% in controls); 40% reduction in maternal body weight gain; 40% reduction in the percentage of viable litters; decreased mean birth weight and decreased mean 3-day postnatal weight gain.

Bergman et al. (1990) administered diets containing 0.15, 3.0 or 4.5 ppm of selenium (sodium selenite) to female Sprague-Dawley rats for 8 weeks prior to mating, during mating and up to gestation day 14. Sodium selenite did not induce toxic effects in either the dam or fetuses.

3.1.5. Reference Dose

3.1.5.1. Subchronic

ORAL RfD: 0.005 mg/kg/day (U.S. EPA, 1992c)

UNCERTAINTY FACTOR: 3

NOAEL: 0.015 mg/kg/day, humans (average body weight 55 kg).

3.1.5.2. Chronic

ORAL RfD: 0.005 mg/kg/day (U.S. EPA, 1992a)

UNCERTAINTY FACTOR: 3

NOAEL: 0.015 mg/kg/day, humans (average body weight 55 kg).

LOAEL: 0.023 mg/kg/day, humans.

CONFIDENCE:

Study: Medium

Data Base: High

RfD High

VERIFICATION DATE: 3/27/91

PRINCIPAL STUDIES: Yang et al., 1989a, 1989b

COMMENT: Based on an epidemiologic study of selenosis in humans (U.S. EPA, 1992a). NOAEL calculated from regression analysis based upon correlation between dietary selenium intake and blood selenium level related to clinical selenosis in adults (average body weight 55 kg). The same chronic oral RfD of 0.005 mg/kg/day has been verified for selenious acid (U.S. EPA, 1992b).

3.2. INHALATION EXPOSURES

Selenium compounds most likely to occur in air include elemental selenium and selenium dioxide dusts and hydrogen selenide gas (ATSDR, 1989). Other volatile selenium compounds that might be present are dimethyl selenide and dimethyl diselenide.

3.2.1. Acute Toxicity

3.2.1.1. Human

Inhalation of selenium or selenium compounds affects primarily the respiratory system. Dusts of elemental selenium and selenium dioxide can cause irritation of the mucous membranes of the nose and throat, coughing, nosebleed, loss of sense of smell, dyspnea, bronchial spasms, bronchitis and chemical pneumonia (Clinton, 1947; Hamilton and Hardy, 1949). Concentrations of 0.007-0.05 mg Se/m3 produced tracheobronchitis in 9 of 62 exposed workers (Kinnigkeit, 1962). Hydrogen selenide gas is also a respiratory tract irritant and acute exposures can lead to pulmonary edema, severe bronchitis, and bronchial pneumonia (Buchan, 1947).

Other signs and symptoms following acute inhalation exposures to selenium compounds include lacrimation, irritation and redness of the eyes, gastrointestinal distress (nausea and vomiting), depressed blood pressure, elevated pulse rate, headaches, dizziness, and malaise (ATSDR, 1989).

3.2.1.2. Animal

Exposure of rats for 8 hr to 30 mg Se/m3 (elemental selenium dust) resulted in severe respiratory effects including edema, hemorrhage, and interstitial pneumonitis (Hall et al., 1951). Diffuse bronchopneumonia and pneumonitis occurred in guinea pigs exposed to 7.8 mg Se/m3 for 4 hr (Dudley and Miller, 1937). The 4-hr LC50 was reported to be 9 mg/m3.

Splenic damage (congestion, fissuring red pulp, and increased polymorphonuclear leukocytes) occurred in guinea pigs exposed to 31 Se mg/m3 (elemental selenium dust, 4 hr/day for 8 days); and to the liver (mild congestion and mild central atrophy) in rats exposed to 30 mg Se (dust)/m3 for 8 hr (Hall et al., 1951).

3.2.2. Subchronic Toxicity

Information on the toxicity of selenium or selenium compounds in humans or animals following subchronic inhalation exposures was not available.

3.2.3. Chronic Toxicity

3.2.3.1. Human

Diskin et al. (1979) reported a case study of a 71-year old man occupationally exposed to selenium for 50 years who suffered from an acute myocardial infarction. A chest X-ray showed bilateral basilar infiltrates indicative of congestive heart failure. Postmortem findings showed evidence of generalized coronary atherosclerosis; severe passive congestion of the lungs, spleen, and liver; and numerous perivascular noncaseating granulomas of the lung and some areas of fibrosis. Abnormally high levels of selenium were found in the peribronchial nodes, lung, hair, and nails and moderately high levels in the thyroid and kidney. The levels of selenium in the liver were within the normal range.

3.2.3.2. Animal

Information on the toxicity of selenium or selenium compounds in animals following chronic inhalation exposures was not available.

3.2.4. Developmental and Reproductive Toxicity

Information on the developmental and reproductive toxicity of selenium or selenium compounds in humans or animals following inhalation exposures was not available.

3.2.5. Reference Concentration

3.2.5.1. Subchronic

A subchronic inhalation Reference Concentration for selenium and selenium compounds is not available.

3.2.5.2 Chronic

A chronic inhalation Reference Concentration for selenium and selenium compounds is not available.

3.3. OTHER ROUTES OF EXPOSURE

3.3.1. Acute Toxicity

3.3.1.1. Human

Dermal exposure to dusts of elemental selenium or selenium dioxide can result in skin rashes and contact dermatitis (Clinton, 1947; Glover, 1967; Middleton, 1947; Pringle, 1942).

3.3.1.2. Animal

Information on the acute toxicity of selenium or selenium compounds in animals following dermal exposure or other routes of exposure was not available.

3.3.2. Subchronic Toxicity

Information on the subchronic toxicity of selenium or selenium compounds in humans or animals by dermal or other routes of exposures was not available.

3.3.3. Chronic Toxicity

Information on the chronic toxicity of selenium or selenium compounds in humans or animals by dermal or other routes of exposures was not available.

3.3.4. Developmental and Reproductive Toxicity

3.3.4.1. Human

Information on the developmental and reproductive toxicity of selenium or selenium compounds in humans by dermal or other routes of exposure was not available.

3.3.4.2. Animal

Subcutaneous or intravenous injection of selenium into pregnant mice or hamsters produced no teratogenic effects even at dose levels lethal to the dam (Lee et al., 1979; Yonemoto et al., 1984; Holmberg and Fern, 1969).

Significant testicular degeneration and atrophy was observed in male mice receiving daily intraperitoneal injections of 0.035 mg Se/kg (as selenium dioxide) for 90 days (Chowdhury and Venkatakrisna-Bhatt, 1983).

Ostadalova and Babicky (1980) reported a dose-related increase in eye cataracts in 10-day-old rats receiving single subcutaneous doses of sodium selenate, DL-selenomethionine, or DL-selenocystine. Dimethyl selenide and trimethylselenonium chloride did not cause cataracts.

3.4. TARGET ORGANS/CRITICAL EFFECTS

3.4.1. Oral Exposures

3.4.1.1. Primary Target Organ(s)

1. Nervous system: Peripheral anesthesia, acroparesthesia, pain in the extremities, hyperreflexia of the tendon, numbness, convulsions, paralysis, motor disturbances, and hemiplegia in humans chronically exposed. Listlessness, lack of mental alertness, irritability, chills and tremors following acute and subchronic exposures.

2. Skin and hair: Dermatitis, brittle nails, loss of hair and nails, mottled and eroded teeth in humans chronically exposed.

3. Developmental and reproductive effects: Decreased rates of conception and increased rates of fetal resorption in animals.

4. Liver: Severe cirrhosis and necrosis following chronic exposures in animals. Abnormal liver function in humans following acute exposures.

3.4.1.2. Other Target Organ(s)

1. GI tract: Nausea, vomiting, abdominal pain in humans following acute exposures.

2. Immune system: Reduced humoral antibody (IgC) production and reduced prostaglandin synthesis in animals following subchronic exposures.

3. Blood chemistry: Reduced hemoglobin levels in humans chronically exposed.

3.4.2. Inhalation Exposures

3.4.2.1. Primary Target Organ(s) (chronic exposure data not available)

1. Respiratory tract: Irritation, dyspnea, bronchial spasms, bronchitis and chemical pneumonia in humans and animals following acute exposures.

2. GI tract: Gastrointestinal distress (nausea and vomiting) in humans following acute exposures.

3. CNS: Headaches, dizziness, and malaise in humans following acute exposures.

3.4.2.2. Other Target Organ(s)

1. Liver: Mild congestion and mild central atrophy in animals following acute exposures; severe cirrhosis and necrosis following chronic exposures.

2. Spleen: Congestion, fissuring red pulp, and increased polymorphonuclear leukocytes in animals following acute exposures. No chronic exposure data.

4. CARCINOGENICITY

4.1. ORAL EXPOSURES

4.1.1. Human

Epidemiologic studies in humans have failed to show a positive association between selenium exposure and an increased incidence of death due to neoplasms. The preponderance of data indicates that selenium has anti-neoplastic effects (see Whanger, 1983; Hocman, 1988). Several epidemiologic studies have revealed negative correlations between selenium intake [based on direct or indirect data on consumption (i.e., soil or plant concentration)] or selenium blood levels (based on direct clinical measurement) and cancer incidence or mortality rates (Shamberger and Willis, 1971; Shamberger et al., 1976; Stampfer et al., 1987; Shamberger et al., 1973; Broghamer et al., 1976; Willett et al., 1983). The anti-neoplastic activity of selenium is thought to be due to one or more mechanisms including: (1) activation of glutathione peroxidase, a enzyme that protects cells against oxidative damage caused by superoxide and hydroxide radicals; (2) interference with enzymes that convert carcinogens to active metabolites; (3) enhancement of immunological responses; (4) inhibition of the interaction of carcinogens with DNA; (5) reductions in body weight; and (6) inhibition of cell proliferation (Hocman, 1988).

4.1.2. Animal

Studies on laboratory animals have not demonstrated a correlation between exposure to selenites or selenates and increased tumor incidence. Although Schroeder and Mitchener (1971a) reported a high incidence rate for tumors (62.5% vs 30.8% in controls) and malignant tumors (41.7% vs. 16.9% in controls) in rats dosed with up to 0.42 mg Se/kg/day (as sodium selenite and sodium selenate in drinking water for a lifetime), according to ATSDR (1989), the study was flawed because the treated rats lived longer than the controls. By adjusting the data for differences in life span, no significant differences in tumor incidence rates were measurable (ATSDR, 1989).

Volgarev and Tscherkes (1967) reported tumor incidence rates of 43% (10/23) and 26% (5/19) in male rats exposed for 25 months to dietary equivalents of 0.22 mg Se/kg/day and 0.43 mg Se/kg/day, respectively; however, no controls were used in these tests and in a third experiment no increase in tumor incidence was observed in 100 animals maintained for 25 months on a diet containing the equivalent of 0.22 mg Se/kg/day.

Harr et al. (1967) evaluated the development of histopathologic lesions in Wistar rats maintained for their entire lifetime (32 months) on diets containing 0, 0.5, 2.0, 8.0 or 16.0 ppm Se (as sodium selenite). No neoplastic lesions could be attributed to the exposure to selenium.

Schroeder and Mitchener (1972) conducted a long-term study in which Charles River CD mice were given drinking water containing 3 ppm Se (as sodium selenite). The incidence of malignant tumors in treated mice (15%) was not significantly different from that in the controls (8%).

Jacobs and Forst (1981a) conducted a 2-year study in which Sprague-Dawley rats were given drinking water containing 4 ppm Se (as sodium selenite). No neoplastic lesions were reported.

Although selenite and selenate compounds have not been found to be carcinogenic in laboratory animals, one study has demonstrated that selenium sulfide when administered orally to rats and mice can result in significant increases in the incidence of hepatocellular carcinomas (male and female rats and female mice) and alveolar/bronchiolar carcinomas and adenomas (female mice only) (NCI, 1980a; U.S. EPA, 1992d). A mixture of selenium monosulfide and selenium disulfide (in 0.5% aqueous carboxymethyl cellulose) was administered by gavage at 3 and 15 mg/kg/day to F344 rats (50/sex/group) and at 20 or 100 mg/kg/day to B6C3F1 mice, 7 days/week for 103 weeks. The incidence of hepatocellular carcinomas was 0/50, 0/50 and 14/49 in control, low-dose and high-dose male rats; 0/50, 0/50 and 21/50 in control, low-dose, and high-dose female rats; and 0/49, 1/50 and 22/49 in control, low-dose and high-dose female mice. The incidence of alveolar/bronchiolar carcinomas and adenomas was 2/50, 0/49, 3/50, and 12/49 in the untreated controls, vehicle controls, low-dose and high-dose female mice.

4.2. INHALATION EXPOSURES

Pertinent data regarding the potential carcinogenicity of selenium by the inhalation route were not located in the available literature. There are no epidemiological data to support a causal association between inhalation exposures and induction of cancer in humans, and no long-term animal studies have been conducted.

4.3. OTHER ROUTES OF EXPOSURE

Information on the carcinogenicity of selenium, selenite or selenate compounds in humans or animals by other routes of exposure was not available. The potential carcinogenicity of selenium sulfide by the dermal route has been evaluated in ICR Swiss mice (NCI, 1980b and 1980c). In one study a suspension of selenium sulfide in 0.5% aqueous carboxymethyl cellulose was applied to the clipped backs of the test animals at 0, 0.5 or 1.0 mg/animal three times per week for 86 weeks (NCI, 1980b). The incidence of alveolar/bronchiolar carcinomas and adenomas (16%) in the high-dose females was significantly greater than the vehicle controls (4%), but lower than the untreated controls (18%). The incidence of total hemangiomas or hemangiosarcomas was significantly elevated in female mice (1/50, 0/50, 1/50 and 4/50 for untreated, vehicle control, low-dose and high-dose groups). In a second study, ICR Swiss mice were treated with selenium sulfide-based Selsun shampoo (0.5 mL of 25% or 50% in distilled water) using the same protocol described above. The incidence of alveolar/bronchiolar carcinomas and adenomas was 1/49, 7/50, and 9/48 in the vehicle controls, low- and high-dose groups. Tumors of this type are common in aged Swiss mice and therefore a causal association with exposure to selenium sulfide could not be clearly established (U.S. EPA, 1992d).

4.4. EPA WEIGHT-OF-EVIDENCE

Selenium and selenious acid

Classification -- D; not classifiable as to human carcinogenicity (U.S. EPA, 1992a, 1992b)

Basis -- Inadequate human data and inadequate evidence of carcinogenicity in animals.

Selenium sulfide

Classification -- B2, probable human carcinogen (U.S. EPA, 1992d)

Basis -- Based on inadequate human data and sufficient evidence of carcinogenicity in animals.

4.5. CARCINOGENICITY SLOPE FACTORS

The calculation of slope factors for elemental selenium, selenites, and selenates is not possible due to the lack of evidence of carcinogenicity. The calculation of a slope factor for selenium sulfide is not possible because of lack of suitable quantitative data (U.S. EPA, 1992d).

5. REFERENCES

Anderson, H.D. and A.L. Moxon. 1942. Changes in the blood picture of the dog following subcutaneous injections. J. Pharmacol. Exp. Ther. 76: 343-352. (Cited in ATSDR, 1989).

ATSDR (Agency for Toxic Substances and Disease Registry). 1989. Toxicological Profile for Selenium. Agency for Toxic Substances and Disease Registry, U.S. Public Health Service, Atlanta GA.

Beems, R.B. and L. van Beek. 1985. Short-term (6-week) oral toxicity study of selenium in Syrian hamsters. Fd. Chem. Toxicol. 23: 945-947.

Bergman, K., E. Cekan, P. Slanina, J. Gabrielsson and K.-E. Hellenas. 1990. Effects of dietary sodium selenite supplementation on salicylate-induced embryo- and fetotoxicity in the rat. Toxicology. 61: 135-146.

Bopp, B.A., R.C. Sonders and J.W. Kesterson. 1982. Metabolic fate of selected selenium compounds in laboratory animals and man. Drug. Metab. Rev. 13: 271-318.

Broghamer, W.L.,Jr., K.P. McConnell and A.L. Blotcky. 1976. Relationship between serum selenium levels and patients with carcinoma. Cancer 37: 1384-1388.

Buchan, R.F. 1947. Industrial selenosis. Occup. Med. 3: 439-456. (Cited in ATSDR, 1989).

Burk, R.F., D.G. Brown, R.J. Seely and C.C. Scaief III. 1972. Influence of dietary and injected selenium on whole-body retention, route of excretion, and tissue retention of 75SeO32- in the rat. J. Nutr. 102: 1049-1055.

Butler, J.A., P.D. Whanger, A.J. Kaneps and N.M. Patton. 1990. Metabolism of selenite and selenomethionine in Rhesus monkey. J. Nutr. 120: 751-759.

Carter, R.F. 1966. Acute selenium poisoning. Med. J. Aust. 1: 525-528. (Cited in ATSDR, 1989).

CDC (Centers for Disease Control). 1984. Selenium intoxication - New York. U.S. Department of Health and Human Services, Public Health Services. Morbidity and Mortality Weekly Report. 33: 157-158.

Chowdhury, A.R. and H. Venkatakrishna-Bhatt. 1983. Effect of selenium dioxide on the testes of rats. Indian J. Physiol. Pharmacol. 27: 237-240. (Cited in ATSDR, 1989).

Civil, I.E. and M.J. McDonald. 1978. Acute selenium poisoning: Case report. N. Zealand Med. J. 87: 354-356. (Cited in ATSDR, 1989).

Clinton, M. Jr. 1947. Selenium fume exposure. J. Ind. Hyg. Toxicol. 29: 225-226. (Cited in ATSDR, 1989).

Cummins, L.M. and E.T. Kimura. 1971. Safety evaluation of selenium sulfiide antidandruff shampoos. Toxicol. Appl. Pharmacol. 20: 89-96. (Cited in ATSDR, 1989).

Diskin, C.J., C.L. Tomasso, J.C. Alper, M.L. Glaser and S.E. Fliegel. 1979. Long-term selenium exposure. Arch. Intern. Med. 139: 824-826.

Dudley, H.C. and J.W. Miller. 1937. Toxicology of selenium. IV. Effects of exposure to hydrogen selenide. Public Health Report 52: 1217-1231. (Cited in ATSDR, 1989).

Echevarria M.G., P.R. Henry, C.B. Ammerman and P.V. Rao. 1988. Effects of time and dietary selenium concentration as sodium selenite on tissue selenium uptake by sheep. J. Anim. Sci. 66: 2299-2305.

Furchner, J.E., J.E. London and J.S. Wilson. 1975. Comparative metabolism of radionuclides in mammals - IX. Retention of 75Se in the mouse, rat, monkey and dog. Health Physics. 29: 641-648.

Ganther, H.E. 1979. Metabolism of hydrogen selenide and methylated selenides. Adv. Nutr. Res. 2: 107-128.

Ganther, H.E. and C.A. Baumann. 1962. Selenium metabolism. II. Modifying effects of sulfate. J. Nutr. 77: 408-414.

Glover, J.R. 1967. Selenium in human urine: A tentative maximum allowable concentration for industrial and rural populations. Ann. Occup. Hyg. 10: 3-10. (Cited in ATSDR, 1989).

Goehring, R.B., I.S. Palmer, O.E. Olson, et al. 1984. Toxic effects of selenium on growing swine fed corn-soybean meal diets. J. Anim. Sci. 59: 733-737. (Cited in ATSDR, 1989).

Hall, R.H., S. Laskin, P. Frank, et al. 1951. Preliminary observations on toxicity of elemental selenium. AMA Arch. Ind. Hyg. Assoc. 4: 458-464. (Cited in ATSDR, 1989).

Halverson, A.W., I.S. Palmer and P.L. Guss. 1966. Toxicity of selenium to post-weanling rats. Toxicol. Appl. Pharmacol. 9: 477-484.

Hamilton, A. and H.L. Hardy. 1949. Selenium in industrial toxicology. Hoeber, Inc. New York. pp. 188-192. (Cited in ATSDR, 1989).

Hammond, P.B. and R.P. Beliles. 1980. Metals. In: Casarett and Doull's Toxicology, The Basic Science of Poisons, 2nd ed., J. Doull, C.D. Klaassen, and M.O. Amdur, eds. Macmillam Publ., New York. pp. 409-467.

Hardin, B.D., R.L. Schuler, J.R. Burg, et al. 1987. Evaluation of 60 chemicals in a preliminary developmental toxicity test. Teratog. Carcinog. Mutag. 7: 29-48.

Harr, J.R., J.F. Bone, I.J. Tinsley, P.H. Weswig and R.S. Yamamoto. 1967. Selenium toxicity in rats. II. Histopathology. In: Selenium in Biomedicine, O.H. Muth, ed., First Int. Symp., Oregon State University, 1966. Avi Publ., Westport, CT. pp. 153-178.

Harr, J.R. and O.H. Muth. 1972. Selenium poisoning in domestic animals and its relationship to man. Clin. Toxicol. 5: 175-186. (Cited in ATSDR, 1989).

Harrison, L.H., B.M. Colvin, B.P. Stuart, et al. 1983. Paralysis in swine due to focal symmetrical poliomalacia: Possible selenium toxicosis. Vet. Pathol. 20: 265-273. (Cited in ATSDR, 1989).

Hocman, G. 1988. Chemoprevention of cancer: Selenium. Int. J. Biochem. 20: 123-132.

Holmberg, R.E. and V.H. Ferm. 1969. Interrelationships of selenium, cadmium, and arsenic in mammalian teratogenesis. Arch. Environ. Health 18: 873-877. (Cited in ATSDR, 1989).

Hopper, S.A., A. Greig and C.H. McMurray. 1985. Selenium poisoning in lambs. Vet. Rec. 116: 569-571. (Cited in ATSDR, 1989).

Jacobs, M. and C. Forst. 1981a. Toxicological effects of sodium selenite in Sprague-Dawley rats. J. Toxicol. Environ. Health 8: 575-585.

Jacobs, M. and C. Forst. 1981b. Toxicological effects of sodium selenite in Swiss mice. J. Toxicol. Environ. Health 8: 587-598.

Jiang, S., H. Robberecht and D. Vanden Berghe. 1983. Elimination of selenium compounds by mice through formation of different volatile selenides. Experientia. 39: 293-294.

Julius, A.D., M.H. Davies and D.F. Birt. 1983. Toxic effects of dietary selenium in the Syrian hamster. Ann. Nutr. Metab. 27: 296-305.

Kinnigkeit, G. 1962. Investigation of workers exposed to selenium in a factory producing rectifiers. Bull. Hyg. 37: 1029-1039. [Abstract]. (Cited in ATSDR, 1989).

Koller, L.D., J.H. Exon, P.A. Talcott, et al. 1986. Immune response in rats supplemented with selenium. Clin. Exp. Immunol. 63: 570-576. (Cited in ATSDR, 1989).

Koppel, C., H. Baudisch, K.-H. Beyer, et al. 1986. Fatal poisoning with selenium dioxide. Clin. Toxicol. 24: 21-35. (Cited in ATSDR, 1989).

Lee, M., K.K.-S. Chan, E. Sairenji, et al. 1979. Effect of sodium selenite on methylmercury-induced cleft palate in the mouse. Environ. Res. 19: 39-48. (Cited in ATSDR, 1989).

Longnecker, M.P., P.R. Taylor, O.A. Levander, et al. 1991. Selenium in diet, blood, and toenails in relation to human health in a seleniferous area. Am. J. Clin. Nutr. 53: 1288-1294.

Maag, D.D., J.S. Orsborn and J.R. Clopton. 1960. The effect of sodium selenite on cattle. Am. J. Vet. Res. 21: 1049-1053. (Cited in ATSDR, 1989).

Medinsky, M.A., R.G. Cuddihy and R.O. McClellan. 1981. Systemic absorption of selenious acid and elemental selenium aerosols in rats. J. Toxicol. Environ. Health. 8: 917-928. (Cited in U.S. EPA, 1989).

Middleton, J.M. 1947. Selenium burn of the eye. Review of a case with review of the literature. Arch. Ophthalmol. 38: 806-811. (Cited in ATSDR, 1989).

Nakamuro, K., Y. Sayato and Y. Ose. 1977. Studies on selenium-related compounds VI. Biosynthesis of dimethyl selenide in rat liver after oral administration of sodium selenate. Toxicol. Appl. Pharmacol. 39: 521-529.

NCI (National Cancer Institute). 1980a. Bioassay of Selenium Sulfide (Gavage) for Possible Carcinogenicity. Technical Report Series No. 194, NTP-80-17. National Cancer Institute, National Institutes of Health, Bethesda, MD.

NCI (National Cancer Institute). 1980b. Bioassay of Selenium Sulfide (Dermal Study) for Possible Carcinogenicity. Technical Report Series No. 197, NTP-80-18. National Cancer Institute, National Institutes of Health, Bethesda, MD.

NCI (National Cancer Institute). 1980c. Bioassay of Selsun (Dermal Study) for Possible Carcinogenicity. Technical Report Series No. 199, NTP-80-19. National Cancer Institute, National Institutes of Health, Bethesda, MD.

NRC (National Research Council). 1989. Recommended Dietary Allowances. 10th ed., National Academy Press, Washington, D.C. pp. 217-224.

Nobunaga, T., H. Satch, T. Suzuki. 1979. Effects of sodium selenite on methylmercury embryotoxicity and teratogenicity in mice. Toxicol. Appl. Pharmacol. 47: 79-88.

Ostadalova, I. and A. Babicky. 1980. Toxic effect of various selenium compounds on the rat in the early postnatal period. Arch. Toxicol. 45: 207-211. (Cited in ATSDR, 1989).

Palmer, I.S., D.D. Fischer, A.W. Halverson and O.E. Olson. 1969. [Title not given]. Biochem. Biophys. Acta. 177: 336. (Cited in Bopp et al., 1982)

Palmer, I.S., R.P. Gunsalus, A.W. Halverson and O.E. Olson. 1970. Trimethylselenonium ion as a general excretory product from selenium metabolism in the rat. Biochim. Biophys. Acta. 208: 260-266.

Pletnikova, I.P. 1970. Biological effect and safe concentration of selenium in drinking water. Hyg. Sanit. 35: 176-180. (Cited in ATSDR, 1989).

Pringle, P. 1942. Occupational dermatitis following exposure to inorganic selenium compounds. Brit. J. Derm. Syphil. 54: 54-58. (Cited in ATSDR, 1989).

Rosenfeld, I. and O.A. Beath. 1964. Selenium: Geobotany, biochemistry, toxicity, and nutrition. Academic Press, NY. (Cited in ATSDR).

Schroeder, H.A. 1967. Effects of selenate, selenite and tellurite on the growth and early survival of mice and rats. J. Nutr. 92: 334-338.

Schroeder, H.A. and M. Mitchner. 1971a. Selenium and tellurium in rats: Effect on growth, survival, and tumors. J. Nutr. 101: 1531-1540.

Schroeder, H.A. and M. Mitchener. 1971b. Toxic effects of trace elements on reproduction of mice and rats. Arch. Environ. Health 23: 102-106.

Schroeder, H.A. and M. Mitchener. 1972. Selenium and tellurium in mice: Effects on growth, survival, and tumors. Arch. Environ. Health 24: 66-71.

Shamberger, R.J. 1984. Selenium. In: Biochemistry of the Essential Ultratrace Elements, E. Frieden, ed. Plenum Press, NY. pp. 201-237.

Shamberger, R.J., E. Rukovena, A.K. Longfield, S.A. Tytko, S. Deadhar and C.E. Willis. 1973. Antioxidants and cancer. I. Selenium in the blood of normals and cancer patients. J. Natl. Cancer. Inst. 50: 863-870.

Shamberger, R.J., S.A. Tytko and C.E. Willis. 1976. Antioxidants and cancer. Part VI. Selenium and age-adjusted human cancer mortality. Arch. Environ. Health 31: 231-235.

Shamberger, R.J. and C.E. Willis. 1971. Selenium distribution and human cancer mortality. CRC Crit. Rev. Clin. Lab. Sci. 2: 211-221.

Smith, M.I., E.F. Stohlman and R.D. Lillie. 1937. The toxicity and pathology of selenium. J. Pharmacol. Exp. Ther. 60: 449-471. (Cited in ATSDR, 1989).

Stampfer, M.J., G.A. Colditz and W.C. Willett. 1987. The epidemiology of selenium and cancer. Cancer Surv. 6: 623-633.

Thomson, C.D. 1974. [Title not given]. N. Zealand Med. J. 80: 163. (Cited in Bopp et al., 1982).

Thomson C.D., C.E. Burton and R.D.H. Stewart. 1978. On supplementing the selenium intake of New Zealanders. 1. Short experiments with large doses of selenite or selenomethionine. Br. J. Nutr. 39: 579-587. (Cited in ATSDR, 1989).

Thomson C.D. and R.D.H. Stewart. 1974. The metabolism of [75Se] in young women. Br. J. Nutr. 32: 47-57. (Cited in Bopp et al., 1982; cited in U.S. EPA, 1989)

Tinsley, I.J., J.R. Harr, J.F. Bone, P.H. Weswig and R.S. Yamamoto. 1967. Selenium toxicity in rats. I. Growth and longevity. In: Selenium in Biomedicine, O.H. Muth, ed., First Int. Symp., Oregon State University, 1966. Avi Publishing Co., Inc.: Westport, CT. pp. 141-152.

U.S. EPA. 1986. Reference Values for Risk Assessment. U.S Environmental Protection Agency, Office of Health and Environmental Assessment, Environmental Criteria and Assessment Office, Cincinnati, OH.

U.S. EPA. 1989. Health and Environmental Effects Document for Selenium and Compounds. Final Draft. Office of Health and Environmental Assessment, Environmental Criteria and Assessment Office, Cincinnati, OH. ECAO-CIN-G058.

U.S. EPA. 1992a. Integrated Risk Information System (IRIS). Oral RfD Assessment and Carcinogenicity Assessment for Selenium and Compounds. (Online file). Office of Health and Environmental Assessment, Cincinnati, OH.

U.S. EPA. 1992b. Integrated Risk Information System (IRIS). Oral RfD Assessment and Carcinogenicity Assessment for Selenious Acid. (Online file). Office of Health and Environmental Assessment, Cincinnati, OH.

U.S. EPA. 1992c. Health Effects Assessment Summary Tables. FY 1992 Annual. U.S. Environmental Protection Agency, Office of Solid Waste and Emergency Response, Washington, DC.

U.S. EPA. 1992d. Integrated Risk Information System (IRIS). Carcinogenicity Assessment for Selenium sulfide. (Online file). Office of Health and Environmental Assessment, Cincinnati, OH.

Volgarev, M.N. and L.A. Tscherkes. 1967. Further studies in tissue changes associated with sodium selenate. In: Selenium in Biomedicine, O.H. Muth, ed., First Int. Symp., Oregon Stage University, 1966. Avi Publ. Co., Inc., Westport CT. pp. 179-184.

Wahlstrom, R.C. and O.E. Olson. 1959. The effect of selenium on reproduction in swine. J. Anim. Sci. 18: 141-145. (Cited in ATSDR, 1989).

Weissman, S.H., R.G. Cuddihy and M.A. Medinsky. 1983. Absorption, distribution and retention of inhaled selenious acid and selenium metal aerosols in beagle dogs. Toxicol. Appl. Pharmacol. 67: 331-337. (Cited in U.S. EPA, 1989).

Whanger, P.D. 1983. Selenium interactions with carcinogens. Fund. Appl. Toxicol. 3: 424-430.

Whanger, P.D. 1989. China, a country with both a selenium deficiency and toxicity: Some thoughts and impressions. J. Nutr. 119: 1236-1239.

Whiting, R.F., L. Wei and H.F. Stich. 1980. Unscheduled DNA synthesis and chromosome aberrations induced by inorganic and organic selenium compounds in the presence of glutathione. Mutat. Res. 78: 159-169.

Willett, W.C., B.F. Polk and J.S. Morris, et al. 1983. Prediagnostic serum selenium and risk of cancer. Lancet. II: 130-134.

Wilson, T.M. and T.R. Drake. 1982. Porcine focal symmetrical poliomyelomalacia. Can. J. Comp. Med. 46: 218-220. (Cited in ATSDR, 1989).

Wilson, T.M., R.W. Scholz and T.R. Drake. 1983. Selenium toxicity and porcine focal symmetrical poliomyelomalacia: Description of a field outbreak and experimental reproduction. Can. J. Comp. Med. 47: 412-421. (Cited in ATSDR, 1989).

Yang, G., S. Wang, R. Zhou, and S. Sun. 1983. Endemic selenium intoxication of humans in China. Am. J. Clin. Nutr. 37: 872-881.

Yang, G., R.Zhou, S. Yin, et al. 1989a. Studies of safe maximal daily dietary selenium intake in a seleniferous area in China. I. Selenium intake and tissue selenium levels of the inhabitants. J. Trace Elem. Electrolyte Health Dis. 3: 77-87.

Yang, G., S. Yin, R. Zhou, et al. 1989b. Studies of safe maximal daily dietary Se-intake in a seleniferous area in China. II. Relation between Se-intake and the manifestation of clinical signs and certain biochemical alterations in blood and urine. J. Trace Elem. Electrolyte Health Dis. 3: 123-130.

Yonemoto, J., T. Hongo, T. Suzuki, et al. 1984. Toxic effects of selenodiglutathione on pregnant mice. Toxicol. Lett. 21: 35-39. (Cited in ATSDR, 1989).

Retrieve Toxicity Profiles Condensed Version

Last Updated 2/13/98

For information or technical assistance, please contact Fred Dolislager.

 

Last updated on Wednesday, August 17th, 2005
URL: http://rais.ornl.gov/tox/profiles/selenium_f_V1.shtml