National Academies Press: OpenBook
« Previous: 3 Ethylene Glycol
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 126
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 127
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 128
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 129
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 130
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 131
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 132
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 133
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 134
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 135
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 136
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 137
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 138
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 139
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 140
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 141
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 142
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 143
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 144
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 145
Suggested Citation:"4 Methanol." National Research Council. 2008. Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3. Washington, DC: The National Academies Press. doi: 10.17226/12527.
×
Page 146

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 Methanol Hector D. García, Ph.D. Toxicology Group Habitability and Environmental Factors Division Johnson Space Center National Aeronautics and Space Administration Houston, Texas PHYSICAL AND CHEMICAL PROPERTIES Methanol, the simplest alcohol, is a colorless, volatile, highly flammable liquid with a mild, characteristic, agreeable odor. Its physical and chemical properties are shown in Table 4-1. TABLE 4-1 Physical and Chemical Properties of Methanol Formula CH3OH Chemical name Methanol Synonyms Methyl alcohol, wood spirit, carbinol, Columbian spirits, wood alcohol, wood naphtha, methyl hydroxide, hydroxy- methane, methyl hydrate, methylol, monohydroxymethane CAS registry no. 67-56-1 Molecular weight 32.04 Density (H2O = 1) 0.79 Boiling point 65°C Melting point –98°C Vapor pressure 12.3 kPa at 20°C Explosive limits 5.5% to 44% in air Solubility Completely miscible in water, ethanol, ether, and many other organic solvents Conversion factor 1 part per million (ppm) = 1.31 milligrams per cubic meter (mg/m3) 126

Methanol 127 OCCURRENCE AND USE Methanol occurs naturally in humans, animals, and plants, including fresh fruits and vegetables, and in fermented products, including wine and other spir- its (see Tables 4-2 and 4-3). It is produced from the distillation of wood and is synthesized catalytically from crude petroleum. It is used industrially in the manufacture of other chemicals and as a solvent. It is added to various commer- cial and consumer products, including windshield washing solutions, deicing solutions, glass cleaners, duplicating fluids, solid canned fuels, paint thinners and removers, model airplane fuels, embalming fluids, lacquers, inks, and some formulations of gasohol motor fuel. Methanol has been found at concentrations up to 3.7 milligrams per liter (mg/L), with an average of <0.1 mg/L, in drinking water on the International Space Station (M. Hwang, J. Schultz, M. Hwang, and J. Schultz, National Aeronautics and Space Administration, Houston, TX, per- sonal commun., 2005). TOXICOKINETICS AND METABOLISM In the early 1980s, measurements of the background concentration of methanol in human blood used available techniques that were limited by poor sensitivity (limit of detection = about 0.4 milligram per deciliter [mg/dL]) TABLE 4-2 Methanol Concentrations in Foods and Beverages Beverages and Food Concentration Beverages Fresh and canned fruit juices Average of 140 mg/L (orange and grapefruit juices) 1-43 mg/L 11-80 mg/L 12-640 mg/L Carbonated beverages ~56 mg/L Neutral spirits <1,500 mg/L Beer 6-27 mg/L Wines 96-329 mg/L Distilled spirits 16-220 mg/L Bourbon 55 mg/L Grain alcohol (50%) 1 mg/L Brandies (USA, Canada, and Italy) 6,000-7,000 mg/L Food Beans 1.5-7.9 mg/kg Split peas 3.6 mg/kg Lentils 4.4 mg/kg Abbreviation: mg/kg, milligram per kilogram. Source: NTP CERHR 2003.

128 Spacecraft Water Exposure Guidelines TABLE 4-3 Background Blood Methanol and Formate Concentrations in Humans Methanol, mg/dL Formate, mg/dL Mean ± SD Mean ± SD Subjects (Range) (Range) Reference Twelve males on restricted diet 0.0570 ± 0.0305 0.38 ± 0.11 Cook et al. (no methanol-containing or (0.025-0.14) (0.22-0. 66) 1991 methanol-producing foods), 12 h Twenty-two adults on restricted 0.18 ± 0.26 1.12 ± 0.91 Chuwers et al. diet (no methanol-containing or (no range data) (no range data) 1995, Osterloh methanol-producing foods), 24 h et al. 1996 Three males who ate a breakfast 0.182 ± 0.121 0.908 ± 0.126 Lee et al. 1992 with no aspartame-containing (0.057-0.357) (0.731-1.057) cereals and no juice Five males who ate a breakfast 0.193 ± 0.093 0.878 ± 0.182 Lee et al. 1992 with no aspartame-containing (0.054-0.315) (0.536-1.083) cereals and no juice. (Second experiment) Twelve adults who drank no 0.18 ± 0.07 No data Batterman and alcohol for 24 h (no range data) Franzblau 1997 Twelve adults who drank no 0.17 ± 0.09 No data Batterman et alcohol for 24 h (0.04-0.47) al. 1998 Thirty adults who fasted for 10 h <0.4 1.91 Stegink et al. (no range data) (no range data) 1981 Twenty-four infants who fasted <0.35 No data Stegink et al. for an unspecified time (no range data) 1983 Abbreviation: mg/dL, milligram per deciliter. Source: NTP CERHR 2003. (Stegink et al. 1981). As shown in Table 4-3, later (1992-1998) measurements with a superior limit of detection of about 0.05 mg/dL consistently showed background concentrations of methanol in blood of about 0.18 mg/dL in adults on methanol-restricted diets (Lee et al. 1992; Chuwers et al. 1995; Osterloh et al. 1996; Batterman and Franzblau 1997; Batterman et al. 1998). Absorption Gastrointestinal absorption of methanol is rapid and almost total after oral exposure. Inhalation studies with humans have shown a net absorption of methanol of 60% to 85% (NTP CERHR 2003). In autopsies of 28 humans who ingested lethal quantities of methanol, however, four had methanol in their stomach contents several days after the initial intoxication, possibly because it

Methanol 129 distributes in all body water, thus being resecreted to gastric fluids after the ini- tial absorption (Mittal et al. 1991). In pregnant monkeys exposed for 2.5 h/d, 7 d/wk to 0, 200, 600, and 1,800 ppm of methanol vapor, average blood methanol concentrations 30 min postex- posure were 5, 11, and 35 micrograms per milliliter (µg/mL) (0.5, 1.1, and 3.5 mg/dL) in the groups exposed to 200, 600, and 1,800 ppm, respectively (Burba- cher et al. 2004a). Blood concentrations in the 200-ppm group were barely above the basal (preexposure) blood methanol concentrations or those observed in the control group (about 3 µg/ml; 0.3 mg/dL). Short-term inhalation exposure of adult volunteers to methanol vapors at 800 ppm for 0, 0.5, 1, 2, and 8 h produced maximal average blood methanol concentrations of 0.06, 0.53, 0.66, 1.40, and 3.07 mg/dL, respectively (Batter- man et al. 1998). Distribution Absorbed methanol is distributed rapidly (peak blood concentrations were reached 30-90 min postexposure (Nashed and Fink 1994)) and uniformly to all organs and tissues in proportion to their water content (NTP CERHR 2003). In humans, the overall volume of distribution is 0.5-0.6 L/kg body weight (Liesi- vuori and Savolainen 1991; IPCS 1997). Thus, a 70-kg person would have 35-42 L of water in all tissues. Because microgravity causes fluid shifts to which the body responds by increasing the excretion of water into urine and, over a period of a few days after entering microgravity, results in a reduced blood volume, the lower value (35 L) of the range should apply to astronauts. Metabolism Humans and animals metabolize methanol primarily in the liver by se- quential oxidative steps, first to formaldehyde by alcohol dehydrogenase (ADH) in humans and by catalase in rodents. Ethanol can act as a competitive inhibitor of ADH and is used therapeutically to slow the metabolism of methanol. For- maldehyde is rapidly oxidized (half-life, about 1 min) in humans and nonhuman primates to formic acid. In a tetrahydrofolate-dependent reaction, primates slowly metabolize formic acid to carbon dioxide, but nonprimates metabolize it about twice as rapidly (McMartin et al. 1977). The delayed toxicity of ingested methanol in humans appears to be due to the accumulation of the toxic metabo- lite, formate. The major route of formate metabolism in the rat, monkey, and presumably humans is oxidation to carbon dioxide via the folate biochemical pathway. Formate enters the pathway by combining with tetrahydrofolate. The susceptibility of various species to methanol toxicity is inversely related to the rate of tetrahydrofolate-dependent oxidation of formate to carbon dioxide (Black et al. 1985). Overall metabolism in humans proceeds with reported half-times of

130 Spacecraft Water Exposure Guidelines 24 h or more with doses greater than 1 g/kg and half-times of 2.5-3 h for doses less than 0.1 g/kg (IPCS 1997). Two mechanisms may be operative in explaining low formate oxidation in species susceptible to methanol toxicity: low hepatic tetrahydrofolate concentra- tions and reduced hepatic 10-formyltetrahydrofolate dehydrogenase activity (Johlin et al. 1987). The activity of 10-formyltetrahydrofolate dehydrogenase, the enzyme that catalyzes the final step of formate oxidation to carbon dioxide, has been reported to be markedly (4- to 5-fold) reduced in monkey and human liver compared with rodent liver, consistent with lower rates of formate oxida- tion in primates. Both mechanisms should apply to formate generated by me- tabolism of ingested methanol. Formate did not accumulate in the blood of volunteers exposed to metha- nol vapor at 200 ppm for 6 h (Lee et al. 1992), either at rest or with light exer- cise. The limit of detection for formate in that study was about 0.3 mg/dL. Blood concentrations of methanol increased from 1.8 µg/mL (0.18 mg/dL) before ex- posure to 7.0 µg/mL (0.70 mg/dL) after 6 h at rest and 8.1 µg/mL (0.81 mg/dL) after 6 h with light exercise. Excretion If there is no simultaneous exposure to ethanol, the kidneys excrete 2% to 5% of ingested methanol unchanged and the lungs eliminate a small amount. At blood concentrations below 15 mg/L (15 µg/ml; 1.5 mg/dL), the half-life of in- haled methanol in the blood of human volunteers has been reported to be 1.44 ± 0.33 h (Batterman et al. 1998), but, at blood concentrations above 300 mg/dL (3,000 µg/mL)—that is, after ingestion of a toxic dose—the enzymes that me- tabolize methanol become saturated, the half-life increases to 27 h (Tephly 1991), and the kidneys and lungs eliminate a larger proportion of unchanged methanol . Simultaneous exposure to ethanol (a therapeutic measure at a target ethanol concentration of 100-150 mg/dL in blood) increases the half-life of methanol to 30-52 h. TOXICITY SUMMARY Selected studies of various toxic effects of methanol are summarized in Table 4-4. Most of the available literature on methanol toxicity in humans con- cerns the effects of acute rather than chronic exposures and shows interindi- vidual variability in the doses that produce toxic effects. Ingested doses as low as 25 mL of 40% methanol (= 7.9 g of methanol) have been reported to be le- thal, but, in other cases, individuals have survived doses up to 500 mL of “moonshine” containing 40% methanol (Kavet and Nauss 1990; Nashed and Fink 1994).

TABLE 4-4 Toxicity Summary Concentration/Dose, Exposure chemical form Route Duration Species/Strain Effects Reference Acute Exposures (<10 d) Unknown high dose Ingestion Single bolus Humans, n = 2 Severe metabolic acidosis, optic disc edema, McMartin et al. death in 1 of 2 1980 Unknown high dose Ingestion Single bolus Humans, n = 2 Severe metabolic acidosis, optic disc edema, Hayreh et al. 1977, blindness 1980 Unknown high dose Ingestion Single bolus Humans, n=3 Severe metabolic acidosis, optic disc edema, Jacobsen and blindness McMartin 1986 192 ppm vapors Inhalation 75 min Humans, n = 12 NOAEL for sensory, neurobehavioral, and Cook et al. 1991 reasoning effects 200 ppm vapors Inhalation 4h Humans, n = 26 NOAEL for visual, neurophysiologic, and Chuwers et al 1995 neurobehavioral effects 200 ppm vapors Inhalation 4h Humans, n = 26 No significant increase in formate in serum or D’Alessandro et al. urine 1994 200 ppm vapors Inhalation 4h Humans, n = 22 Serum methanol increased 4-fold to 6.5 mg/L; Osterloh et al. 1996 half-life of methanol in serum = 3.2 ± 2.3 h 200 ppm vapors Inhalation 6h Humans, n = 6 No increase in formate in blood; serum Lee et al. 1992 methanol increased to 0.7 mg/dL at rest and 0.81g/dL with light exercise 200 ppm vapors Inhalation 2h Humans, n = 4 No increase in urinary formic acid; methanol in Ernstgard et al. 2005 blood increased from 20 to 244 µM 800 ppm, vapors Inhalation 8h Humans, n = 15 Measured methanol in breath, blood, and urine; Batterman et al. no toxic effects were examined 1998 (Continued) 131

TABLE 4-4 Continued 132 Concentration/Dose, Exposure chemical form Route Duration Species/Strain Effects Reference Aspartame, 200 mg/kg Ingestion 3 times / d Humans, n = 30 Blood methanol increased from <0.4 mg/dL to Stegink et al. 1981 (equivalent to 20 mg of a peak of 2.58 mg/dL with no increase in methanol / kg) formate in blood Subchronic Exposures (11-100 d) 2,500 mg of Ingestion 90 d Rats, male and LOAEL for elevated SGPT and SAP and EPA 1986 methanol/kg/d female, Sprague- decreased brain weight Dawley, n = 30/sex/dose 365-3,080 ppm vapors Inhalation 1 h/wk – 40 Teacher aids, Dose-dependent incidence of blurred vision, Frederick et al. 1984 h / wk female, n = 66 headache, dizziness, and nausea Chronic Exposures (≥100 d) 32.7 mg of metha- Ingestion 3 doses/d, 7 Human, n = 53 NOAEL for standard lab tests or symptoms Leon et al. 1989 nol/dose (as metabolite d/wk, 6 mo of 900 mg aspartame/d) Abbreviations: LOAEL, lowest-observed-adverse-effect level; NOAEL, no-observed-adverse-effect level; SAP, serum alkaline phosphatase; SGPT, serum glutamic-pyruvic transaminase (now called alanine amino transferase).

Methanol 133 The initial toxic effect of methanol is inebriation within the first 1-2 h af- ter ingestion. After a delay of 6-30 h following ingestion, more serious toxic effects are associated with methanol: headaches, dizziness, nausea, blurred vi- sion, gastric and muscle pain, metabolic acidosis, blindness due to optic nerve injury, and, in extreme exposures, death. The severity and kinetics of toxic ef- fects depend on the rate of metabolism in the liver, which can vary substantially among individuals, depending on the hepatic concentration of folate, a cofactor needed to convert formate to carbon dioxide, and depending on concurrent ex- posure to (treatment with) substances such as ethanol and fomepizole (4-methyl pyrazole), which can inhibit ADH and slow production of the toxic metabolite, formic acid or formate. Interpretation of toxicity data from animal experiments involving expo- sures to methanol is complicated by interspecies as well as nutritional status- dependent differences in the rate of metabolism of methanol. In rodents, the initial oxidation of methanol to formaldehyde is catalyzed by catalase, rather than by ADH as in humans. Because catalase is saturated at low concentrations of methanol, the blood concentrations of methanol increase in rodents, whereas in humans, the conversion of formate to carbon dioxide is the rate-limiting step in metabolism and formate accumulates in the blood, in addition to methanol. The developmental toxicity (Nelson et al. 1985; Rogers et al. 1993; Weiss et al. 1996) reported in rodents is of doubtful relevance to humans because it occurs only at doses (mg/kg) severalfold above the lethal dose in humans (Methanol Institute 2002). Acute Exposures (≤10 d) Symptoms of acute methanol toxicity begin with narcosis followed by a symptom-free period of several hours to about a day, followed by visual distur- bances and central nervous system (CNS) effects, including headache; dizziness; abdominal, back, and leg pain; nausea; and delirium that can lead to coma (Gosselin et al. 1984). Many of the delayed toxic effects can be attributed to the accumulation of formate, which also leads to metabolic acidosis. Neurobehavioral Effects Potential CNS effects of inhalation of 192 ppm of methanol vapor for 75 min were tested in 12 young male volunteers (Cook et al. 1991). A battery of 20 neurobehavioral tests was administered before, during, and after each exposure. No detectable effects were seen except for two tests in which minor effects were within the range of test values seen in control subjects exposed to air. No significant neurobehavioral effects were seen in 26 volunteers exposed once for 4 h to 200 ppm of methanol vapors and tested for color discrimination (Stroop test and Lanthony 15), contrast sensitivity (Visitech), auditory evoked

134 Spacecraft Water Exposure Guidelines potentials, visual scanning, symbol-digit substitution, memory scanning (Stern- berg), and event-related potential (P-300), unless certain between-subject vari- ables were considered (Chuwers et al. 1995). Specifically, significant effects were seen on only two outcomes: the P-300 amplitude when alcohol consump- tion and smoking accounted for between-subject variability (P = 0.01) and the symbol-digit test with age accounting for between-subject variability (P = 0.05). In agreement with previous reports (D’Alessandro et al. 1994), formate concen- trations in serum were not affected after exposure to 200 ppm of methanol for 4 h, but serum methanol concentrations were increased (peak = 0.65 mg/dL, which was 0.56 mg/dL greater than in controls). CNS Degeneration The sensitivity of the CNS to degenerative changes after exposure to methanol is supported by a case report of putaminal necrosis seen in computed tomography scans of the head of a 26-y-old woman with degenerating mental status after recovery from acute methanol intoxication (Fontenot and Pelak 2002). An autopsy study of 28 fatal cases of methanol poisoning, in which shrinkage and degeneration of neurons in the parietal cortex were seen in 86% of the cases and putaminal degeneration and necrosis were seen in 7.1% of the cases, provided further evidence of the CNS toxicity of methanol (Mittal et al. 1991). Ocular Effects Ocular toxicity is commonly seen in cases of human exposure to metha- nol. McMartin et al. reported severe metabolic acidosis and optic disc edema in two patients whose blood concentrations were 11.1 and 26.0 milliequivalents of formate/mL when measured after hospitalization for methanol intoxication (McMartin et al. 1980). The patient with the higher blood formate concentration died despite aggressive treatment (bicarbonate, intravenous 10% ethanol, and hemodialysis) of both patients. In humans, permanent visual damage has been associated with prolonged exposures (>24 h) to blood formate concentrations >7 mM (31.5 mg/dL) pro- duced during methanol intoxication (Hayreh et al. 1980; Hayreh et al. 1977; Jacobsen and McMartin 1986). Clinical observations in methanol-intoxicated humans have shown that, in its initial stages, developing ocular toxicity can be reversed, even >24 h after methanol ingestion, by treatments such as bicarbon- ate, fomepizole or ethanol, intravenous folinic acid, and dialysis, which reduce blood formate concentrations and metabolic acidosis (Barceloux et al. 2002). Such reversibility is consistent with a mechanism of formate toxicity involving gradual optic nerve toxicity due to prolonged metabolic hypoxia and with the reversibility of formate inhibition of cytochrome c oxidase (Nicholls 1975).

Methanol 135 Subchronic and Chronic Exposures (≥10 d) A National Institute for Occupational Safety and Health (NIOSH) epide- miology study of adverse health effects in 66 female teacher aides exposed to methanol vapors from duplicating fluid (99% methanol) found that aides ex- posed to 365-3,080 ppm of methanol reported significantly more blurred vision, headache, dizziness, and nausea than a comparison group of 66 unexposed fe- male teachers randomly selected from 287 respondents to the questionnaire (Frederick et al. 1984). Exposure durations ranged from 1 h/d, 1 d/wk to 8 h/d, 5 d/wk. Fifteen of 21 measurements exceeded the (1984 relevant) NIOSH- recommended 15-min exposure limit of 800 ppm. The NIOSH recommended exposure limit (REL) is currently 200 ppm, time-weighted average (TWA). The incidence of symptoms related to methanol toxicity was positively correlated with the percentage of time spent per week at the duplicator. Aides who spent an average of 85% of their time (based on a 40-h wk) at or near a duplicating ma- chine had a 50% incidence of reported methanol toxicity, which Frederick et al. defined as having any one of four aggregations of symptoms from a list chosen by an independent NIOSH physician. Only about 15% of aides who spent about 5% of their time (about 2 h) near a duplicator reported symptoms of methanol toxicity. The minimum and maximum daily doses the aides potentially inhaled can be estimated from the given ranges of atmospheric concentrations and expo- sure durations: Minimum: 365 ppm × 1.31 mg/m3/ppm × 20 m3/24 h × 60% × 1 h/d = 240 mg/d. Maximum: 3,080 ppm × 1.31 mg/m3/ppm × 20 m3/24 h × 60% × 8 h/d = 16,000 mg/d. An earlier study (Kingsley and Hirsch 1955) reported frequent headaches and dizziness, but no blurred vision, in workers exposed to methanol vapors at concentrations of 15-375 ppm (20-490 mg/m3). In a recent study, serum metha- nol concentrations in volunteers inhaling 200 ppm of methanol for 4 h were 0.65 ± 0.27 mg/dL, and serum methanol concentrations in controls were 0.09 ± 0.06 mg/dL (Osterloh et al. 1996). A case report was found attributing progressive parkinsonism in a young experimental physicist to long-term exposure to methanol vapors in the labora- tory, with no episodes of acute intoxication (Finkelstein and Vardi 2002). No estimate was provided of the dose or exposure concentration of methanol, and it appears that there was simultaneous exposure to methyl bromide. A study of the safety of 6 mo of exposure to 300 mg of aspartame (an arti- ficial sweetener that is rapidly and completely metabolized to methanol and its constituent amino acids) three times/d (equivalent to 32.7 mg of methanol/dose or 98.1 mg of methanol/d) reported no elevation in blood methanol concentra-

136 Spacecraft Water Exposure Guidelines tion (most concentrations were below the level of detection, 0.992 mg/dL), no significant increases in blood formate or mean urinary formate concentrations, and no adverse effects associated with the exposure (Leon et al. 1989). The Environmental Protection Agency (EPA)-sponsored 90-d gavage study in rats, tested at methanol concentrations of 0, 100, 500, and 2,500 mg/kg of body weight/d (10 rats/sex/dose group), found increases in liver enzymes (serum glutamic pyruvic transaminase (SGPT) and serum alkaline phosphatase (SAP)) and decreases in brain weight at the highest dose but no histopathologic lesions in the liver (EPA 1986). A dose rate of 500 mg/kg/d was considered a no-observed-adverse-effect level (NOAEL) in rats. No consistent treatment-related effects were found for organ or body weights or for histopathologic or ophthalmoscopic examinations in a subchronic inhalation study of rats and monkeys exposed to 0, 500, 2,000, and 5,000 ppm of methanol for 6 h/d, 5 d/wk for 4 wk (Andrews et al. 1987). Genotoxicity Methanol was negative for cell transformation in Syrian hamster embryo cells (clonal assay and viral enhanced), for sister chromatid exchange in vitro, and for aneuploidy and chromosome aberrations in Neurospora crassa (EPA 1994). The micronucleus test and the assay for chromosome aberrations in mammalian polychromatic erythrocytes were inconclusive (EPA 1994). Reproductive Toxicity No studies of the potential reproductive toxicity of methanol in humans were found. In male Sprague-Dawley rats, inhalation of 200 ppm of methanol for 8 h/d, 5 d/wk for 6 wk reduced serum testosterone concentrations to 32% that of controls, although similar exposure to 10,000 ppm of methanol had no significant effect on testosterone concentration but increased serum levels of luteinizing hormone about 3-fold (Cameron et al. 1984). In female monkeys, exposures for 2.5 h/d, 7 d/wk to 0, 200, 600, and 1,800 ppm of methanol vapor did not alter their menstrual cycles, the number of breedings to conception or the conception rate, or the birth weight or health status of the offspring but de- creased the length of pregnancy by 6 to 8 d compared with controls (Burbacher et al. 2004b). Developmental and Fetal Toxicity No studies were found of the potential developmental or fetal toxicity of methanol in humans. In CD-1 mice, inhaled methanol vapors caused develop- mental abnormalities (neural and ocular defects, cleft palate, hydronephrosis,

Methanol 137 deformed tails, and limb (paw and digit) anomalies) at doses of 10,000 and 15, 000 ppm for 6 h/d, 3 d (Bolon et al. 1993). The highest concentration was mildly to moderately neurotoxic to up to 20% of the dams but would yield a dose that would be potentially lethal in humans (about 98 g/d). Similar results were re- ported in an EPA study of CD-1 mice, with a NOAEL of 1,000 ppm, 7 h/d on days 6-15 of gestation (Rogers et al. 1993). Plasma methanol concentrations in this study were approximately 9.7, 53.7, 165.0, 317.8, 420.4, and 733.0 mg/dL in the 1,000-, 2,000-, 5,000-, 7,500-, 10,000-, and 15,000-ppm exposure groups, respectively. In pregnant rats and their newborn pups, exposure to 4,500 ppm of methanol vapor produced only isolated, subtle performance decrements in a bat- tery of neurobehavioral tests (Weiss et al. 1996). Further testing is needed to establish whether the observed differences are reproducible and significant. Offspring of rats consuming 2% methanol (an average of 2.5 g/kg of body weight) in drinking water during gestational days 15-17 or 17-19 required longer than controls consuming distilled water to begin suckling on postnatal day 1; they also required more time than the controls to locate nesting material from their home cages on postnatal day 10 (Infurna and Weiss 1986). No other overt toxicity was observed in dams or pups. Neither the Clean Air Act (CAA), Clean Water Act (CWA), nor the Safe Drinking Water Act (SDWA) require methanol monitoring. Some current stan- dards for methanol exposures for workers and the general public are listed in Tables 4-5 and 4-6. Basis for EPA’s Reference Dose Calculation In an EPA-sponsored study, Sprague-Dawley rats that were gavaged daily with methanol at 0, 100, 500, or 2,500 mg/kg/d for 90 d showed elevated SGPT and SAP concentrations and small, but not statistically significant, increases in liver weight. Brain weights were significantly decreased in males and females at the highest dose. A dose rate of 500 mg/kg/d was considered a NOAEL in rats. An oral reference dose (RfD) of 0.5 mg/kg/d was calculated from the NOAEL by applying uncertainty factors of 10 for interindividual variations, 10 for spe- cies extrapolation, and 10 for subchronic to chronic extrapolation. However, rodents are poor models for methanol toxicity in humans. In addition, for a given daily dose rate, the use of gavage treatments would be expected to result TABLE 4-5 Drinking Water Standards for Methanol Set by Other Organizations Organization, Reference Standard Amount Concentration EPA 1993 Reference dose 0.5 mg/kg/d 12.5 mg/La (oral, chronic) a The calculated concentration is based on a body weight of 70 kg and 2.8 L of water/d.

138 Spacecraft Water Exposure Guidelines TABLE 4-6 Air Standards for Methanol Vapors Set by Other Organizations Organization, Standard Amount Reference NIOSH NIOSH 2005 IDLH 6,000 ppm REL-TWA 200 ppm (260 mg/m3) ST 250 ppm (325 mg/m3) (skin) OSHA NIOSH 2005 PEL-TWA 200 ppm (260 mg/m3) ACGIH ACGIH 2008 TLV-TWA 200 ppm, skin TLV-STEL 250 ppm, skin Abbreviations: ACGIH, American Conference of Governmental Industrial Hygienists; IDLH, immediately dangerous to life and health; NIOSH, National Institute for Occupa- tional Safety and Health; OSHA, Occupational Safety and Health Administration; PEL, permissible exposure limit; REL, recommended exposure limit; ST, short-term exposure limit (15 min); STEL, short-term exposure limit; TLV, Threshold Limit Value; TWA; time-weighted average. in higher peak blood methanol concentrations compared with what would be expected from smaller fractionated doses throughout the day for drinking water exposures. RfD values are meant to be protective of the entire population, in- cluding the elderly, children, and unusually sensitive individuals. NIOSH does not state the basis of its REL value of 200 ppm but based its immediately-dangerous-to-life-and-health value of 6,000 ppm on a Russian re- port of a 2-h LCLo (the lowest tested concentration that killed any mice after two hours of exposure) in mice of 37,594 ppm (Izmerov et al. 1982). The Occupational Safety and Health Administration does not state the ba- sis of its value of 200 ppm for the permissible exposure limit (PEL), but it often uses the Threshold Limit Values (TLVs) of the American Conference of Gov- ernmental Industrial Hygienists (ACGIH) in setting PEL values. ACGIH did not state the specific studies or reports used as the basis for setting a TLV of 200 ppm but states that “Based on the human and animal toxi- cological responses, the TLV Committee recommends a TLV-TWA of 200 ppm and a STEL [short-term exposure limit] of 250 ppm for methanol” (ACGIH 1991). Since exposure scenarios for spacecraft crews differ significantly from those of workers or the general public, NASA calculates exposure guidelines for specific exposure durations on the basis of the lowest acceptable concentration for the various adverse effects as listed in Table 4-7 and described in the section below.

Methanol 139 TABLE 4-7 Spacecraft Water Exposure Guidelines for Methanol Duration Concentration, mg/L Target Toxicity 1d 40 Neurobehavioral effects 10 d 40 Neurobehavioral effects 100 d 40 Neurobehavioral effects 1,000 d 40 Neurobehavioral effects RATIONALE Acceptable concentrations (ACs) (Table 4-8) were determined following the guidelines of the National Research Council (NRC 2000). ACs were calcu- lated as described below assuming consumption of 2.8 L of water/d. This in- cludes an average of 800 mL/d used to prepare or reconstitute food in addition to 2.0 L/d for drinking. A value of 70 kg was used as the mass of an average astronaut. Spaceflight Effects None of the reported adverse effects of methanol exposures are known to be affected by spaceflight, but the 1.7% to 12% reduction in total body water associated with prolonged microgravity would proportionately increase the blood concentration of any methanol ingested. This effect is too small to signifi- cantly change the spacecraft water exposure guideline (SWEG) calculations. One-Day SWEG Immediate and delayed adverse effects of acute exposures to toxic, but not life-threatening, doses of methanol, such as those reported by teacher’s aids and other workers intermittently exposed to methanol vapors, include narcosis, headaches, dizziness, nausea, and blurred vision. More subtle neurobehavioral effects in humans (e.g., slight effects on P-300 amplitude and symbol-digit sub- stitution tests) have been detected in controlled exposures to low (200 ppm) concentrations of methanol vapors. The available evidence suggests that acute exposures that produce such effects but do not produce symptoms of metabolic acidosis are not usually associated with permanent sequelae. Calculation of a concentration of methanol in drinking water that will keep blood methanol concentrations from accumulating to toxic levels when exposure is intermittent requires knowing the half-life of methanol in the blood. The pub- lished data on the half-life of methanol at low doses are sparse—only four pa- pers (Haffner et al. 1992, 1997; Osterloh et al. 1996; Batterman et al. 1998), in addition to the study of Chuwers et al. (1995), were found, involving a total of

140 Spacecraft Water Exposure Guidelines 28 healthy subjects. The reported range of half-lives in individuals ranged from 1.2 to 3.6 h. A similar narrow range of half-lives was found in reviews of the literature for the metabolism of ethanol, despite the existence of marked genetic polymorphisms in some of the ADH enzymes, which metabolize both methanol and ethanol. This masking of the genetic variability may be explained by the contributions of multiple ADH enzymes and CYP2E1 to alcohol elimination. Note that the longer half-lives (e.g., 20 h) for blood methanol that have been reported occur only at concentrations above 300 mg/dL, such as those produced after ingestion of large doses of methanol. Thus, in calculating an acceptable concentration of methanol in drinking water, the use of the longest half-life (3.6 h) seen in any tested individual at low doses should provide a high level of con- fidence that methanol concentrations that do not exceed the calculated concen- trations will not produce methanol toxicity in astronauts. Because Chuwers’ study involved fewer than 100 volunteers, the observed NOAEL of 0.65 mg/dL = 650 µg/dL should be adjusted by a factor of √n/√100 = √26/√100 = 0.51 to account for potential interindividual variations (NRC 2000). Thus, the target peak concentration of blood methanol should be 0.65 mg/dL × √26/√100 = 0.33 mg/dL. The maximum allowable concentration of methanol in drinking water that will produce a concentration of 0.33 mg/dL in the blood can be calculated as follows. If the half-life of methanol in the blood at low concentrations is ≤3.6 h, we can calculate the time course of methanol concentration in the blood during a typical day when drinking water contains a specified low concentration of methanol. Astronauts consume about 2.8 L of water per day. The pattern of con- sumption varies somewhat from individual to individual and from day to day, but for the purposes of this calculation we can assume that the 2.8 L is con- sumed in volumes equivalent to that of a canned soft drink (about 670 mL, in- cluding water used to reconstitute food) at meal times (assume 8 AM, noon, and 7 PM) with additional drinks of about 400 mL at around 3 PM and 9 PM. This pattern of drinking water consumption was used previously in calculating SWEG values for formate. Plots of the blood methanol concentration at each hour during a typical day can be used to determine the maximum concentration of methanol in drinking water that will maintain a blood methanol concentration of ≤0.33 mg/dL. Figure 4-1 shows that drinking water containing methanol at 40 mg/L will maintain blood methanol concentrations below 330 µg/dL of blood when the half-life is 3.6 h. The hourly calculated concentration of methanol in the blood is shown in Figure 4-1.

TABLE 4-8 Acceptable Concentrations for Methanol Uncertainty Factor Acceptable Concentrations, mg/L Exposure Inter- LOAEL to Inter- Space- End Point Data Species individual NOAEL species Time flight 1d 10 d 100 d 1,000 d NOAEL for neuro- 200 ppm, Human, 0.51 1 1 PK model 1 40 40 40 40 behavioral toxicity 4h n = 26 SWEG 40 40 40 40 141

142 Spacecraft Water Exposure Guidelines 1 1-Day Blood Methanol Kinetics 0.9 For fractionated consumption of 2.8 L of water/day containing 0.8 40 mg methanol/L, assuming a 3.6-h half-life mg Methanol/dL Blood 0.7 0.6 670 ml 400 ml 670 ml 670 ml 400 ml 0.5 0.4 0.3 0.2 0.1 0 0 4 8 12 16 20 24 Hours FIGURE 4-1 Calculated 24-h blood methanol kinetics for a 70-kg person ingesting wa- ter containing methanol at 40 mg/L. The pattern of ingestion was assumed to be 670 mL at 8 h (e.g., 8:00 AM), 670 mL at 12 h, 400 mL at 15 h, 670 mL at 19 h, and 400 mL at 21 h. Chuwers et al. measured a baseline of blood methanol at 0.18 mg/dL (Chuwers et al. 1995). Ten-, One-Hundred-, and One-Thousand-Day SWEGs Because the toxic effects of methanol do not follow Haber’s rule but in- stead depend on the concentration of methanol (or its metabolite, formate) in the blood, SWEG values for exposures >1 d should be set to ensure that methanol does not accumulate in the blood—that is, the peak methanol concentration does not increase during a multiday exposure. Thus, the methodology used to calcu- late blood methanol kinetics for a 1-d SWEG of 40 mg/L can also be applied to calculating SWEG values for 10, 100, and 1,000 d. Extending the exposure to examine the blood methanol concentration over multiple days reveals that the peak methanol concentration does not increase from day to day, but each day it achieves the same peak concentration and remains below the calculated target NOAEL concentration of 0.33 mg/dL (Figure 4-2). This pattern repeats for all exposure durations of ≥1 d.

Methanol 143 1 0.9 5-Day Blood Methanol Kinetics For fractionated consumption of 2.8 L of water/day containing 0.8 40 mg methanol/L, assuming a 3.6-h half-life mg Methanol/dL Blood 0.7 0.6 670 ml 670 ml 400 ml 670 ml 670 ml 400 ml 670 ml 670 ml 400 ml 670 ml 670 ml 400 ml 670 ml 670 ml 400 ml 670 ml 400 ml 670 ml 400 ml 670 ml 400 ml 400 ml 670 ml 400 ml 670 ml 0.5 0.4 0.3 0.2 0.1 0 0 4 8 12 16 20 24 28 32 36 40 44 48 52 56 60 64 68 72 76 80 84 88 92 96 100 104 108 112 116 120 124 128 Hours FIGURE 4-2 Calculated multiday blood methanol kinetics for a 70-kg person ingesting water containing methanol at 40 mg/L. The pattern of ingestion was assumed to be 670 mL at 8 h (e.g., 8:00 AM), 670 mL at 12 h, 400 mL at 15 h, 670 mL at 19 h, and 400 mL at 21 h, repeating each 24-h period. REFERENCES ACGIH (American Conference of Governmental Industrial Hygienists). 1991. Methanol. Pp. 903-905 in Documentation of the Threshold Limit Values and Biological Ex- posure Indices, 6th Ed. Cincinnati, OH: American Conference of Governmental Industrial Hygienists. ACGIH (American Conference of Governmental Industrial Hygienists). 2008. Methanol. Pp. 113 in Guide to Occupational Exposure Values. Cincinnati, OH: American Conference of Governmental Industrial Hygienists. Andrews, L.S., J.J. Clary, J.B. Terrill, and H.F. Bolte. 1987. Subchronic inhalation toxic- ity of methanol. J. Toxicol. Environ. Health 20(1-2):117-124. Barceloux, D.G., G.R. Bond, E.P. Krenzelok, H. Cooper, J.A. Vale, and American Acad- emy of Clinical Toxicology Ad Hoc Committee on the Treatment Guidelines for Methanol Poisoning. 2002. American Academy of Clinical Toxicology practice guidelines on the treatment of methanol poisoning. J. Toxicol. Clin. Toxicol. 40(4):415-446. Batterman, S.A., and A. Franzblau. 1997. Time-resolved cutaneous absorption and per- meation rates of methanol in human volunteers. Int. Arch. Occup. Environ. Health 70(5):341-351. Batterman, S.A., A. Franzblau, J.B. D'Arcy, N.E. Sargent, K.B. Gross, and R.M. Schreck. 1998. Breath, urine, and blood measurements as biological exposure indices of

144 Spacecraft Water Exposure Guidelines short-term inhalation exposure to methanol. Int. Arch. Occup. Environ. Health 71(5):325-335. Black, K.A., J.T. Eells, P.E. Noker, C.A. Hawtrey, and T.R. Tephly. 1985. Role of he- patic tetrahydrofolate in the species difference in methanol toxicity. Proc. Nat. Acad. Sci. U.S.A. 82(11):3854-3858. Bolon, B., D.C. Dorman, D. Janszen, K.T. Morgan, and F. Welsch. 1993. Phase-specific developmental toxicity in mice following maternal methanol inhalation. Fundam. Appl. Toxicol. 21(4):508-516. Burbacher, T.M., D.D. Shen, B. Lalovic, K.S. Grant, L. Sheppard, D. Damian, S. Ellis, and N. Liberato. 2004a. Chronic maternal methanol inhalation in nonhuman pri- mates (Macaca facicularis): Exposure and toxicokinetics prior to and during preg- nancy. Neurotoxicol. Teratol. 26(2):201-221. Burbacher, T.M., K.S. Grant, D.D. Shen, L. Sheppard, D. Damian, S. Ellis, and N. Lib- erato. 2004b. Chronic maternal methanol inhalation in nonhuman primates (Macaca fascicularis): Reproductive performance and birth outcome. Neurotoxi- col. Teratol. 26(5):639-650. Cameron, A.M., O.G. Nilsen, E. Haug, and K.B. Eik-Nes. 1984. Circulating concentra- tions of testosterone, luteinizing hormone and follicle stimulating hormone in male rats after inhalation of methanol. Arch. Toxicol. (Suppl. 7):441-443. Chuwers, P., J. Osterloh, T. Kelly, A. d'Alessandro, P. Quinlan, and C. Becker. 1995. Neurobehavioral effects of low-level methanol vapor exposure in healthy human volunteers. Environ. Res. 71(2):141-150. Cook, M.R., F.J. Bergman, H.D. Cohen, M.M. Gerkovich, C. Graham, R.K. Harris, and L.G. Siemann. 1991. Effects of Methanol Vapor on Human Neurobehavioral Measures. Research Report No. 42. Health Effect Institute, Boston, MA. d'Alessandro, A., J. Osterloh, P. Chuwers, P. Quinlan, T. Kelly, and C. Becker. 1994. Formate in serum and urine after controlled methanol exposure at the threshold limit value. Environ. Health Perspect. 102(2):178-181. Eells, J.T., M.M. Salzmann, M.F. Lewandowski, and T.G. Murray. 1996. Formate- induced alterations in retinal function in methanol-intoxicated rats. Toxicol. Appl. Pharmacol. 140(1):58-69. EPA (U.S. Environmental Protection Agency). 1986. Rat Oral Subchronic Toxicity Study with Methanol. Office of Solid Waste, U.S. Environmental Protection Agency, Washington, DC. EPA (U.S. Environmental Protection Agency). 1993. Methanol. Integrated Risk Informa- tion System, U.S. Environmental Protection Agency [online]. Available: http://www.epa.gov/iris/subst/0305.htm [accessed July 25, 2008]. EPA (U.S. Environmental Protection Agency). 1994. Chemical Summary for Methanol. EPA 749-F-94-013a. Office of Pollution Prevention and Toxics, U.S. Environ- mental Protection Agency. August 1994 [online]. Available: http://www.epa.gov/ chemfact/s_methan.txt [accessed July 30, 2008]. Ernstgard, L., E. Shibata, and G. Johanson. 2005. Uptake and disposition of inhaled methanol vapor in humans. Toxicol. Sci. 88(1):30-38. Finkelstein, Y., and J. Vardi. 2002. Progressive parkinsonism in a young experimental physicist following long-term exposure to methanol. Neurotoxicology 23(4- 5):521-525. Fontenot, A.P., and V.S. Pelak. 2002. Development of neurologic symptoms in a 26-year- old woman following recovery from methanol intoxication. Chest 122(4):1436- 1439.

Methanol 145 Frederick, L.J., P.A. Schulte, and A. Apol. 1984. Investigation and control of occupa- tional hazards associated with the use of spirit duplicators. Am. Ind. Hyg. Assoc. J. 45(1):51-55. Gosselin, R.E., R.P. Smith, and H.C. Hodge. 1984. Clinical Toxicology of Commercial Products, 5th ed. Baltimore, Williams and Wilkins. Haffner, H.T., H.D. Wehner, K.D. Scheytt, and K. Besserer. 1992. The elimination kinet- ics of methanol and the influence of ethanol. Int. J. Legal Med. 105(2):111-114. Haffner, H.T., K. Besserer, M. Graw, and S. Voges. 1997. Methanol elimination in non- alcoholics: Inter- and intraindividual variation. Forensic Sci. Int. 86(1-2):69-76. Hayreh, M.S., S.S. Hayreh, G.L. Baumbach, P. Cancilla, G. Martin-Amat, T.R. Tephly, K.E. McMartin, and A.B. Makar. 1977. Methyl alcohol poisoning. III. Ocular tox- icity. Arch. Ophthalmol. 95(10):1851-1858. Hayreh, M.S., S.S. Hayreh, G. Baumbach, P. Cancilla, G. Martin-Amat, and T.R. Tephly. 1980. Ocular toxicity of methanol: An experimental study. Pp. 35-53 in Neurotox- icity of the Visual System, W. Merrigan, and B. Weiss, eds. New York: Lippin- cott-Raven. Infurna, R., and B. Weiss. 1986. Neonatal behavioral toxicity in rats following prenatal exposure to methanol. Teratology 33(3): 259-265. IPCS (International Programme on Chemical Safety). 1997. Methanol. Environmental Health Criteria 196. U.N. Environment Programme, International Labour Organi- sation, World Health Organization [online]. Available: http://www.inchem.org/ documents/ehc/ehc/ehc196.htm#SectionNumber:1.5. [accessed June 1, 2007]. Izmerov, N.F., I.V. Sanotskii, and K.K. Sidorov. 1982. Toxicometric Parameters of In- dustrial Toxic Chemicals under Single Exposure. Moscow: Center of International Projects, GKNT. Jacobsen, D., and K.E. McMartin. 1986. Methanol and ethylene glycol poisonings. Mechanism of toxicity, clinical course, diagnosis and treatment. Med. Toxicol. 1(5):309-334. Johlin, F.C., C.S. Fortman, D.D. Nghiem, and T.R. Tephly. 1987. Studies on the role of folic acid and folate-dependent enzymes in human methanol poisoning. Mol. Pharmacol. 31(5):557-561. Kavet, R., and K.M. Nauss. 1990. The toxicity of inhaled methanol vapors. Crit. Rev. Toxicol. 21(1):21-50. Kingsley, W.H., and F.G. Hirsch. 1955. Toxicologic considerations in direct process spirit duplicating machines. Compen. Med. 40:7-8. Lee, E.W., T.S. Terzo, J.B. D'Arcy, K.B. Gross, and R.M. Schreck. 1992. Lack of blood formate accumulation in humans following exposure to methanol vapor at the cur- rent permissible exposure limit of 200 ppm. Am. Ind. Hyg. Assoc. J. 53(2):99-104. Leon, A.S., D.B. Hunninghake, C. Bell, D.K. Rassin, and T.R. Tephly. 1989. Safety of long-term large doses of aspartame. Arch. Intern. Med. 149(10):2318-2324. Liesivuori, J., and H. Savolainen. 1991. Methanol and formic acid toxicity: Biochemical mechanisms. Pharmacol. Toxicol. 69(3):157-163. McMartin, K.E., G. Martin-Amat, A.B. Makar, and T.R. Tephly. 1977. Methanol poison- ing. V. Role of formate metabolism in the monkey. J. Pharmcol. Exp. Therap. 201(3):564-572. McMartin, K.E., J.J. Ambre, and T.R. Tephly. 1980. Methanol poisoning in human sub- jects. Role for formic acid accumulation in the metabolic acidosis. Am. J. Med. 68(3):414-418.

146 Spacecraft Water Exposure Guidelines Methanol Institute. 2002. Comments on Final Draft, NTP-CERHR Expert Panel Report on Reproductive and Developmental Toxicity of Methanol. Washington, DC: Methanol Institute. Mittal, B.V., A.P. Desai, and K.R. Khade. 1991. Methyl alcohol poisoning: An autopsy study of 28 cases. J. Postgrad. Med. 37(1):9-13. Nashed, A., and G. Fink. 1994. Methanol [online]. Available: http://www.embbs.com/ cr/alc/alc6.html [accessed March 30, 2005]. Nelson, B.K., W.S. Brightwell, D.R. MacKenzie, A. Khan, J.R. Burg, W.W. Weigel, and P.T. Goad. 1985. Teratological assessment of methanol and ethanol at high inhala- tion levels in rats. Fundam. Appl. Toxicol. 5(4):727-736. Nicholls, P. 1975. Formate as an inhibitor of cytochrome c oxidase. Biochem. Biophys. Res. Commun. 67(2):610-616. NIOSH (National Institute for Occupational Safety and Health). 2005. NIOSH Pocket Guide to Chemical Hazards. DHHS (NIOSH) No. 2005-151. National Institute for Occupational Safety and Health, Centers for Disease Control and Prevention, U.S. Department of Health and Human Services, Cincinnati, OH. NRC (National Research Council). 2000. Methods for Developing Spacecraft Water Ex- posure Guidelines. Washington, DC: National Academy Press. NTP-CERHR (National Toxicology Program–Center for the Evaluation of Risks to Hu- man Reproduction). 2003. NTP-CERHR Expert Panel Report on Reproductive and Developmental Toxicity of Methanol. National Toxicology Program-Center for the Evaluation of Risks to Human Reproduction, U.S. Department of Health and Hu- man Services, Research Triangle Park, NC. Osterloh, J.D., A. D'Alessandro, P. Chuwers, H. Mogadeddi, and T.J. Kelly. 1996. Serum concentrations of methanol after inhalation at 200 ppm. J. Occup. Environ. Med. 38(6):571-576. Rogers, I.M., M.L. Mole, N. Chernoff, B.D. Barbee, C.I. Turner, T.R. Logsdon, and R.J. Kavlock. 1993. The developmental toxicity of inhaled methanol in the CD-1 mouse, wiht quantitative dose-response modeling for estimation of benchmark doses. Teratology 47(3):175-188. Stegink, L.D., M.C. Brummel, K.E. McMartin, G. Martin-Amat, L.J. Filer, G.L. Baker, and T.R. Tephly. 1981. Blood methanol concentrations in normal adult subjects administered abuse doses of aspartame. J. Toxicol. Environ. Health 7(2):281-290. Stegink, L.D., M.C. Brummel, L.J. Filer, and G.L. Baker. 1983. Blood methanol concen- trations in one-year-old infants administered graded doses of aspartame. J. Nutr. 113(8):1600-1606. Tephly, T.R. 1991. The toxicity of methanol. Life Sci. 48(11):1031-1041. Weiss, B., S. Stern, S.C. Soderhohn, C. Cox, A. Sharma, G.B. Inglis, R. Preston, M. Ba- lys, K.R. Reuhl, and R. Gelein. 1996. Developmental Neurotoxicity of Methanol Exposure by Inhalation in Rats. Reseach Report No. 73. Health Effects Institute, Boston, MA.

Next: 5 Methyl Ethyl Ketone »
Spacecraft Water Exposure Guidelines for Selected Contaminants: Volume 3 Get This Book
×
Buy Paperback | $62.00 Buy Ebook | $49.99
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

NASA maintains an active interest in the environmental conditions associated with living and working in spacecraft and identifying hazards that might adversely affect the health and well-being of crew members. Despite major engineering advances in controlling the spacecraft environment, some water and air contamination is inevitable. Several hundred chemical species are likely to be found in the closed environment of the spacecraft, and as the frequency, complexity, and duration of human space flight increase, identifying and understanding significant health hazards will become more complicated and more critical for the success of the missions.

To protect space crews from contaminants in potable and hygiene water, NASA requested that the National Research Council NRC provide guidance on how to develop water exposure guidelines and subsequently review NASA's development of the exposure guidelines for specific chemicals. This book presents spacecraft water exposure guidelines (SWEGs) for antimony, benzene, ethylene glycol, methanol, methyl ethyl ketone, and propylene glycol.

  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!