DRAFT - DO NOT CITE OR QUOTE                         EPA/635/R-09/013


                                WWW.EPA.GOV/IRIS
         &EPA
               TOXICOLOGICAL REVIEW

                                     OF

                            METHANOL

                               (CAS No. 67-56-1)


                 In Support of Summary Information on the
                 Integrated Risk Information System (IRIS)

                                October 2009


                                   NOTICE

    This document is an Interagency Science Consultation draft. This information is distributed
    solely for the purpose of pre-dissemination peer review under applicable information quality
    guidelines. It has not been formally disseminated by EPA. It does not represent and should not be
    construed to represent any Agency determination or policy. It is being circulated for review of its
    technical accuracy and science policy implications.

1                           U.S. Environmental Protection Agency
2                                  Washington, DC


                                      1 -1       DRAFT-DO NOT CITE OR QUOTE

-------
                                        DISCLAIMER
1          This document is a preliminary draft for review purposes only. This information is
2   distributed solely for the purpose of pre-dissemination peer review under applicable information
3   quality guidelines. It has not been formally disseminated by EPA. It does not represent and
4   should not be construed to represent any Agency determination or policy. Mention of trade
5   names or commercial products does not constitute endorsement or recommendation for use.
                                              1 -2        DRAFT-DO NOT CITE OR QUOTE

-------
                   CONTENTS TOXICOLOGICAL REVIEW OF METHANOL
                                        (CAS NO. 67-56-1)
 1    1. INTRODUCTION	1-26
 2    2. CHEMICAL AND PHYSICAL INFORMATION	2-1
 3    3. TOXICOKINETICS	3-1
 4      3.1. Overview	3-1
 5      3.2. Key Studies	3-5
 6      3.3. Human Variability In Methanol Metabolism	3-10
 7      3.4. Physiologically Based Toxicokinetic Models	3-15
 8         3.4.1. Model Requirements for EPA Purposes	3-15
 9           3.4. l.l.MOA and Selection of a Dose Metric	3-15
10           3.4.1.2. Criteria for the Development of Methanol PBPK Models	3-17
11         3.4.2. Methanol PBPK Models	3-18
12           3.4.2.1. Ward etal. (1997)	3-19
13           3.4.2.2. Bouchard etal.  (2001)	3-19
14           3.4.2.3. Gentry et al. (2003, 2002) and Clewell et al. (2001)	3-20
15         3.4.3. Selected Modeling Approach	3-21
16           3.4.3.1. Available PK Data	3-22
17           3.4.3.2. Model Structure	3-23
18           3.4.3.3. Model Parameters	3-25
19         3.4.4. Mouse Model Calibration and Sensitivity Analysis	3-26
20         3.4.5. Rat Model Calibration	3-31
21         3.4.6. Human Model Calibration	3-35
22           3.4.6.1. Inhalation Route	3-35
23           3.4.6.2. Oral Route	3-41
24         3.4.7. Monkey PK Data and Analysis	3-42
25           3.4.7.l.PK Model Analysis for Monkeys	3-44
26         3.4.8. Summary and Conclusions	3-47
27    4. HAZARD IDENTIFICATION	4-1
28      4.1. Studies in Humans - Case Reports, Occupational and Controlled Studies	4-1
29         4.1.1. Case Reports	4-1
30         4.1.2. Occupational Studies	4-9
31         4.1.3. Controlled Studies	4-11
32      4.2. Acute, Subchronic and Chronic Studies and Cancer Bioassays in Animals—Oral and
33         Inhalation	4-13
34         4.2.1. Oral Studies	4-13
35           4.2.1.1. Acute Toxicity	4-13
36           4.2.1.2. Subchronic  Toxicity	4-13
37           4.2.1.3. Chronic Toxicity	4-14
38         4.2.2. Inhalation Studies	4-20
39           4.2.2.1. Acute Toxicity	4-20
40           4.2.2.2. Subchronic  Toxicity	4-21
41           4.2.2.3. Chronic Toxicity	4-24
42      4.3. Reproductive AND Developmental Studies-Oral and Inhalation	4-32
43         4.3.1. Oral Studies	4-32
44         4.3.2. Inhalation Studies	4-34
                                               1 -3        DRAFT-DO NOT CITE OR QUOTE

-------
 1         4.3.3. Other Reproductive and Developmental Toxicity Studies	4-46
 2      4.4. Neurotoxicity	4-50
 3         4.4.1. Oral Studies	4-51
 4         4.4.2. Inhalation Studies	4-53
 5         4.4.3. Studies Employing In Vitro, S.C. and IP. Exposures	4-60
 6      4.5. Immunotoxicity	4-64
 7      4.6. Mechanistic Data and Other Studies in Support of the MO A	4-69
 8         4.6.1. Role of Methanol and Metabolites in the Developmental Toxicity of Methanol ..4-69
 9         4.6.2. Role of Folate Deficiency in the Developmental Toxicity of Methanol	4-74
10         4.6.3. Methanol-Induced Formation of Free Radicals, Lipid Peroxidation, and Protein
11            Modifications	4-75
12         4.6.4. Exogenous Formate Dehydrogenase as a Means of Detoxifying the Formic Acid that
13            Results from Methanol Exposure	4-78
14         4.6.5. Mechanistic Data Related to the Potential Carcinogenicity of Methanol	4-78
15            4.6.5.1. Genotoxicity	4-78
16            4.6.5.2. Lymphoma Responses Reported in ERF Life  span Bioassays of Compounds
17                 Related to Methanol, Including an Analogue (Ethanol), Precursors (Aspartame
18                 and Methyl Tertiary Butyl Ether), and a Metabolite (Formaldehyde)	4-81
19            4.6.5.2.1. Ethanol	4-83
20            4.6.5.2.2. Aspartame	4-84
21            4.6.5.2.3. MTBE	4-89
22            4.6.5.2.4. Formaldehyde	4-90
23      4.7. Synthesis of Major Noncancer Effects	4-92
24         4.7.1. Summary of Key Studies in Methanol Toxicity	4-92
25            4.7.1.1. Oral	4-95
26            4.7.1.2. Inhalation	4-96
27      4.8. NONCANCER MOA Information	4-99
28      4.9. Evaluation of Carcinogenicity	4-102
29         4.9.1. Summary of Overall Weight-of-Evidence	4-102
30         4.9.2. Synthesis of Human, Animal, and Other Supporting Evidence	4-103
31         4.9.3. MOA Information	4-109
32      4.10.  Susceptible Populations and Life Stages	4-110
33         4.10.1. Possible Childhood Susceptibility	4-110
34         4.10.2. Possible Gender Differences	4-111
35         4.10.3. Genetic Susceptibility	4-112
36    5. DOSE-RESPONSE ASSESSMENT AND CHARACTERIZATION	5-1
37      5.1.  Inhalation RfC for Effects Other than Cancer	5-1
38         5.1.1. Choice of Principal  Study and Critical Effect(s)	5-1
39            5.1.1.1. Key Inhalation Studies	5-1
40            5.1.1.2. Selection of Critical Effect(s)	5-2
41         5.1.2. Methods of Analysis for the POD—Application of PBPK and BMD Models	5-6
42            5.1.2.1. Applicati on of the BMD/BMDL Approach	5-7
43            5.1.2.2. BMD Approach Applied to Brain Weight Data in Rats	5-9
44         5.1.3. RfC Derivation - Including Application of Uncertainty Factors	5-13
45            5.1.3.1. Comparison Between Endpoints andBMDL Modeling Approaches	5-13
46            5.1.3.2. Applicati on of UFs	5-15
47            5.1.3.2.1. Interindividual Variation UFH	5-15
48            5.1.3.2.2. Animal-to-Human Extrapolation UFA	5-17

                                               1 -4        DRAFT-DO NOT CITE OR QUOTE

-------
 1             5.1.3.2.3. Database UFD	5-17
 2             5.1.3.2.4. Extrapolation from Subchronic to Chronic and LOAEL-to NOAEL
 3             Extrapolation UFs	5-18
 4         5.1.4. Previous RfC Assessment	5-18
 5      5.2. Oral RfD	5-18
 6         5.2.1. Choice of Principal Study and Critical Effect-with Rationale and Justification... 5-19
 7            5.2.1.1. Expansion of the Oral Database by Route-to-Route Extrapolation	5-20
 8         5.2.2. RfD Derivation-Including Application of UFs	5-21
 9            5.2.1.2. Consideration of Inhalation Data	5-21
10            5.2.1.3. Selection of Critical Effect(s) from Inhalation Data	5-21
11            5.2.1.4. Selection of the POD	5-22
12         5.2.2. RfD Derivation-Application of UFs	5-22
13         5.2.3. Previous RfD Assessment	5-22
14      5.3. Uncertanties in the Inhalation RfC and Oral RfD	5-23
15         5.3.1. Choice of Endpoint	5-24
16         5.3.2. Choice of Dose Metric	5-25
17         5.3.3. Choice of Model for BMDL Derivations	5-26
18         5.3.4. Choice of Animal-to-Human Extrapolation Method	5-27
19         5.3.5. Route-to-Route Extrapolation	5-27
20         5.3.6. Statistical Uncertainty at the POD	5-28
21         5.3.7. Choice of Bioassay	5-28
22         5.3.8. Choice of Species/Gender	5-28
23         5.3.9. Human Population Variability	5-28
24      5.4. 5.4. Cancer Assessment	5-29
25         5.4.1. Oral Exposure	5-29
26            5.4.1.1. Choice of Study/Data—with Rationale and Justification	5-29
27            5.4.1.2. Dose-Response Data	5-29
28            5.4.1.3. Dose Adjustments and Extrapolation Method	5-30
29            5.4.1.4. Oral Slope Factor	5-33
30         5.4.2. Inhalation Exposure	5-33
31            5.4.2.1. Choice of Study/Data-with Rationale and Justification	5-33
32            5.4.2.2. Dose-Response Data	5-34
33            5.4.2.3. Dose Adjustments and Extrapolation Method	5-35
34            5.4.2.4. IUR	5-36
35         5.4.3. Uncertainties in Cancer Risk Assessment	5-36
36            5.4.3.1. Quality of Studies that are the Basis for the PODs	5-37
37            5.4.3.2. Interpretation of Results of the Studies that are the Basis for the PODs	5-40
38            5.4.3.3. Consistency across Chronic Bioassays for Methanol	5-43
39            5.4.3.4. Choice of Endpoint for POD Derivation	5-45
40            5.4.3.5. Choice of Species/Gender	5-45
41            5.4.3.6. Choice of Model for POD Derivation	5-45
42            5.4.3.7. Choice of Dose Metric	5-46
43            5.4.3.8. Choice of Animal-to-Human Extrapolation Method	5-47
44            5.4.3.9. Human Relevance of Cancer Responses Observed in Rats and Mice	5-48
45    6. MAJOR CONCLUSIONS IN THE CHARACTERIZATION OF HAZARD AND DOSE
46    RESPONSE	6-1
47      6.1. Human Hazard Potential	6-1
48      6.2. Dose Response	6-3

                                               1 -5        DRAFT-DO NOT CITE OR QUOTE

-------
 1        6.2.1. Noncancer/Inhalation	6-3
 2        6.2.2. Noncancer/Oral	6-5
 3        6.2.3. Cancer/Oral and Inhalation	6-6
 4   7. REFERENCES	7-1
 5   APPENDIX A. SUMMARY OF EXTERNAL PEER REVIEW AND PUBLIC COMMENTS
 6   AND DISPOSITION	A-l
 7   APPENDIX B. DEVELOPMENT, CALIBRATION AND APPLICATION OF A METHANOL
 8   PBPK MODEL	B-l
 9   APPENDIX C. RFC DERIVATION OPTIONS	C-l
10   APPENDIX D. RFC DERIVATION - COMPARISON OF DOSE METRICS	D-l
11   APPENDIX E. EVALUATION OF THE CANCER POTENCY OF METHANOL	E-1
                                      1 -6      DRAFT-DO NOT CITE OR QUOTE

-------
                                        LIST OF TABLES
 1    Table 2-1.  Relevant physical and chemical properties of methanol	2-1
 2    Table 3-1.  Background blood methanol and formate levels in humans	3-11
 3    Table 3-2.  Human blood methanol and formate levels following methanol exposure	3-12
 4    Table 3-3.  Monkey blood methanol and formate levels following methanol exposure	3-13
 5    Table 3-4.  Mouse blood methanol and formate levels following methanol exposure	3-13
 6    Table 3-5.  Rat blood methanol and formate levels following methanol exposure	3-14
 7    Table 3-6.  Plasma formate concentrations in monkeys	3-15
 8    Table 3-7.  Serum folate concentrations in monkeys	3-15
 9    Table 3-8.  Routes of exposure optimized in models - optimized against blood concentration data3-20
10    Table 3-9.  Key methanol kinetic studies for model validation	3-22
11    Table 3-10. Parameters used in the mouse, rat and human PBPK models	3-25
12    Table 3-11. Primate Kms reported in the literature	3-38
13    Table 3-12. Parameter estimate results obtained using acslXtreme to fit all human data using
14             either saturable or first-order metabolism	3-38
15    Table 3-13. Comparison of LLFs for Michaelis-Menten and first-order metabolism	3-40
16    Table 3-14. Monkey group exposure characteristics	3-47
17    Table 4-1. Mortality rate for subjects exposed to methanol-tainted whiskey in relation to their
18             level of acidosisaa	4-2
19    Table 4-2.  Incidence of carcinogenic responses in Sprague-Dawley rats exposed to methanol in
20             drinking water for up to 2 years	4-18
21    Table 4-3.  Incidence of malignant lymphoma responses in Swiss mice exposed to methanol in
22             drinking water for life	4-20
23    Table 4-4. Histopathological changes in tissues of B6C3F1 mice exposed to methanol via
24             inhalation for 18 months	4-28
25    Table 4-5. Histopathological changes in lung and adrenal tissues of F344 rats exposed to
26             methanol via inhalation for 24 months	4-31
27    Table 4-6. Reproductive and developmental toxicity in pregnant Sprague-Dawley rats exposed to
28             methanol via inhalation during gestation	4-36
29    Table 4-7. Reproductive parameters in Sprague-Dawley dams exposed to methanol during
30             pregnancy then allowed to deliver their pups	4-37
31    Table 4-8. Developmental effects in mice after methanol inhalation	4-39
32    Table 4-9. Benchmark doses at two added risk levels	4-40
33    Table 4-10. Reproductive parameters in monkeys exposed via inhalation to methanol during
34             prebreeding, breeding, and pregnancy	4-42
35    Table 4-11. Mean serum levels of testosterone, luteinizing hormone, and corticosterone (± S.D.)
36             in male Sprague-Dawley rats after inhalation of methanol, ethanol, n-propanol or n-
37             butanol at threshold limit values	4-43
38    Table 4-12. Maternal and litter parameters when pregnant female C57BL/6J mice were injected
39             i.p. with methanol	4-46
40    Table 4-13. Reported thresholds concentrations (and author-estimated ranges) for the onset of
41             embryotoxic effects when  rat and mouse conceptuses were incubated in vitro with
42             methanol, formaldehyde, and formate	4-49
43    Table 4-14. Brain weights of rats exposed to methanol vapors during gestation and lactation 4-58
44    Table 4-15. Effect of methanol on Wistar rat acetylcholinesterase activities	4-61
45    Table 4-16. Effect of methanol on neutrophil functions in in vitro and in vivo studies in male
46             Wistar  rats	4-65
                                               1 -7        DRAFT-DO NOT CITE OR QUOTE

-------
 1   Table 4-17. Effect of intraperitoneally injected methanol on total and differential leukocyte
 2            counts and neutrophil function tests in male Wistar rats	4-66
 3   Table 4-18. Effect of methanol exposure on animal weight/organ weight ratios and on cell counts
 4            in primary and secondary lymphoid organs of male Wistar rats	4-68
 5   Table 4-19. The effect of methanol on serum cytokine levels in male Wistar rats	4-69
 6   Table 4-20. Developmental outcome on GD10 following a 6-hour 10,000 ppm (13,104 mg/m3)
 7            methanol inhalation by CD-mice or formate gavage (750 mg/kg) on GD8	4-70
 8   Table 4-21. Summary of ontogeny of relevant enzymes in CD-I mice and humans	4-71
 9   Table 4-22. Dysmorphogenic effect of methanol and formate in neurulating CD-I mouse
10            embryos in culture (GD8)	4-72
11   Table 4-23. Time-dependent effects of methanol administration on serum liver and kidney
12            function, serum ALT, AST, BUN, and creatinine in control and experimental groups
13            of male Wistar rats	4-77
14   Table 4-24. Effect of methanol administration on male Wistar rats on malondialdehyde
15            concentration in the lymphoid organs  of experimental and control groups and the
16            effect of methanol on antioxidants in spleen	4-77
17   Table 4-25. Summary of genotoxicity studies of methanol	4-80
18   Table 4-26. Incidence of carcinogenic responses in Sprague-Dawley rats exposed to ethanol in
19            drinking water for up to 2 years	4-84
20   Table 4-27. Incidence of lymphomas and leukemias in Sprague-Dawley rats exposed to
21            aspartame via the diet	4-85
22   Table 4-28. Incidence of combined dysplastic hyperplasias, papillomas and carcinomas of the
23            pelvis and ureter and of malignant schwannomas in peripheral nerve in Sprague-
24            Dawley rats exposed to aspartame via the diet	4-86
25   Table 4-29. Incidence of tumors in Sprague-Dawley rats exposed to aspartame from GD12 to
26            natural death	4-87
27   Table 4-30. Comparison of the incidence of combined lymphomas and leukemias in female
28            Sprague-Dawley rats exposed to aspartame in feed for a lifetime, either pre- and
29            postnatally or postnatally only	4-88
30   Table 4-31. Incidence of Ley dig cell testicular tumors and combined lymphomas and leukemias
31            in Sprague-Dawley rats  exposed to MTBE via gavage for 104 weeks	4-90
32   Table 4-32. Incidence of hemolymphoreticular neoplasms on Sprague-Dawley rats exposed  to
33            formaldehyde in drinking water for 104  weeks	4-91
34   Table 4-33. Incidence of leukemias in breeder and offspring Sprague-Dawley rats exposed to
35            formaldehyde in drinking water for 104  weeks (TestBT 7005)	4-92
36   Table 4-34. Summary of studies of methanol toxicity in experimental animals (oral)	4-92
37   Table 4-35. Summary of studies of methanol toxicity in experimental animals (inhalation
38            exposure)	4-93
39   Table 5-1. Summary of studies considered most appropriate for use in derivation of an RfC.... 5-5
40   Table 5-2. The EPA PBPK model  estimates of methanol blood levels (AUC) in rats following
41            inhalation exposures	5-10
42   Table 5-3. Comparison of benchmark dose modeling results for decreased brain weight in male
43            rats at 6 weeks of age using modeled AUC of methanol as a dose metric	5-11
44   Table 5-4. Summary of PODs for critical endpoints, application of UFs and conversion to HEC
45            values using BMD and PBPK modeling	5-14
46   Table 5-5. Summary of uncertainties in methanol noncancer risk assessment	5-23
47   Table 5-6. Incidence data for lymphoma, lympho-immunoblastic, and all lymphomas in male and
48            female Sprague-Dawley rats	5-30
49   Table 5-7. BMD results and oral CSF using all lymphoma in male rats	5-33

                                              1 -8         DRAFT-DO NOT CITE OR QUOTE

-------
 1   Table 5-8. Incidence data for tumor responses in male and female F344 rats	5-34
 2   Table 5-9. BMD results and IUR using pheochromocytoma in female rats	5-36
 3   Table 5-10. Summary of uncertainty in the methanol cancer risk assessment	5-37
 4   Table B-l. Parameters used in the mouse and human PBPK models	B-5
 5   Table B-2. Primate kms reported in the literature	B-29
 6   Table B-3. Parameter estimate results obtained using acslXtreme to fit all human data using
 7             either saturable or first-order metabolism	B-31
 8   Table B-4. Comparison of LLF for Michaelis-Menten and first-order metabolism	B-32
 9   Table B-5. PBPK model predicted Cmax and 24-hour AUC for mice and humans exposed to
10             MeOH	B-35
11   Table B-6. Mouse total MeOH metabolic clearance/metabolites produced following inhalation
12             exposures	B-90
13   Table B-7. Human total MeOH metabolic clearance/metabolites produced from inhalation
14             exposures	B-91
15   Table B-8. Human total MeOH metabolic clearance/metabolites produced following oral
16             exposures	B-91
17   Table B-9. Repeated daily oral dosing of humans with MeOH	B-93
18   Table C-l. EPA's  PBPK model estimates of methanol blood levels (AUC) in rats following
19             inhalation exposures	C-3
20   Table C-2. Comparison of BMDiso results for decreased brain weight in male rats at 6 weeks of
21             age using modeled AUC of methanol as a dose metric	C-4
22   Table C-3. Comparison of BMD0s results for decreased brain weight in male rats at 6 weeks of
23             age using modeled AUC of methanol as a dose metric	C-9
24   Table C-4. EPA's PBPK model estimates of methanol blood levels (Cmax) in rats following
25             inhalation exposures	C-l3
26   Table C-5. Comparison of BMDiso results for decreased brain weight in male rats at 8 weeks of
27             age using modeled Cmax of methanol as a dose metric	C-14
28   Table C-6. Comparison of BMDos modeling results for decreased brain weight in male rats at 8
29             weeks of age using modeled Cmax of methanol as a common dose metric	C-20
30   Table C-7. EPA's  PBPK model estimates of methanol blood levels (Cmax) in mice following
31             inhalation exposures	C-24
32   Table C-8. Comparison of BMD modeling results for cervical rib  incidence in mice using
33             modeled Cmax of methanol as a common dose metric	C-25
34   Table C-9. Comparison of BMD modeling results for cervical rib  incidence in mice using
35             modeled Cmax of methanol as a common dose metric	C-33
36   Table C-10. Comparison of BMD modeling results for VDR in female monkeys using  AUC
37             blood methanol  as the dose metric	C-42
38   Table E-l. Calculation of mg/kg-day doses	E-12
39   Table E-2. Incidence for neoplasms considered for dose-response modeling	E-12
40   Table E-3. Results from multistage (1°) quantal modeling rat data using mg/kg-day exposures
41             and default HED derivation method	E-13
42   Table E-4. Results from time-to-tumor modeling data using mg/kg-day exposures and default
43             HED derivation method	E-13
44   Table E-5. PBPK model estimated dose-metrics for doses	E-14
45   Table E-6a. Results from time-to-tumor modeling of data using PBPK dose metrics	E-14
46   Table E-6b. HEDs from time-to-tumor modeling of data using PBPK dose metrics	E-14
47   Table E-7. Results of Multistage (1°) quantal  modeling of data using PBPK dose metrics	E-15
48   Table E-8. Application of human PBPK model to derive HEDs from results of multistage (1°)
49             quantal  modeling of data using PBPK dose metrics	E-15

                                              1 -9        DRAFT-DO NOT CITE OR QUOTE

-------
 1   Table E-9. Incidence for neoplasms considered for dose-response modeling	E-16
 2   Table E-10. PBPK dose metrics for doses	E-16
 3   Table E-l 1. Benchmark results from multistage quantal dose-response modeling data using
 4            PBPK dose-metrics	E-17
 5   Table E-12. Application of human PBPK model to derive HECs from BMDLio estimates in
 6            Table E-ll using multistage quantal modeling	E-17
 7   Table E-13. Incidence for malignant lymphoma (Apaja, 1980)	E-17
 8   Table E-14. PBPK dose metrics for doses in Apaja (1980)	E-18
 9   Table E-l5. Benchmark results from Multistage-cancer dose-response modeling data for
10            malignant lymphoma in Swiss Webster mice (Apaja, 1980) using PBPK dose-metricsE-18
11   Table E-16. Application of human PBPK model to derive HEDs from BMDLio estimates of
12            Table E-15, Multistage (1°) modeling of malignant lymphoma in Swiss mice (Apaja,
13            1980) using PBPK dose metrics	E-19
14   Table E-17. Benchmark results for all tumor types using BMDS 2.1 multistage "background
15            dose" models and PBPK dose-metrics	E-19
                                             1-10        DRAFT-DO NOT CITE OR QUOTE

-------
                                       LIST OF FIGURES

 1   Figure 3-1. Methanol metabolism and key metabolic enzymes in primates and rodents	3-3
 2   Figure 3-2. Folate-dependent formate metabolism. Tetrahydrofolate (THF)-mediated one carbon
 3             metabolism is required for the synthesis of purines, thymidylate, and methionine.... 3-4
 4   Figure 3-3. Plot of fetal (amniotic) versus maternal methanol concentrations in GD20 rats	3-7
 5   Figure 3-4. Schematic of the PBPK model used to describe the inhalation, oral, and i.v. route
 6             pharmacokinetics of methanol	3-24
 7   Figure 3-5. Model fits to data sets from GD6, GD7, and GD10 mice for 6- to 7-hour inhalation
 8             exposures to  1,000-15,000 ppm methanol	3-27
 9   Figure 3-6. Simulation of inhalation exposures to methanol in NP mice from Perkins et al.
10             (1995a) (8-hour exposures) and GD8 mice from Dorman et al. (1995)(6-hour
11             exposures)	3-28
12   Figure 3-7. Conceptus versus maternal blood AUC values for rats and mice plotted (A) on a log-
13             linear scale, as in Figure 8 of Ward et al. (1997), and (B) on a linear-linear scale... 3-30
14   Figure 3-8. NP rat i.v. route methanol blood kinetics	3-33
15   Figure 3-9. Model fits to data sets from inhalation exposures to 200 (triangles), 1,200
16             (diamonds), or 2,000 (squares) ppm methanol in male F-344 rats	3-33
17   Figure 3-10. Model fits to datasets from oral exposures to 100 and 2,500 mg/kg methanol in
18             female Sprague-Dawley rats	3-35
19   Figure 3-11. Urinary methanol elimination concentration (upper panel) and cumulative amount
20             (lower panel) following inhalation exposures to methanol in human volunteers	3-37
21   Figure 3-12. Data showing the visual quality of the fit using optimized first-order or Michaelis-
22             Menten kinetics to describe the metabolism of methanol in humans	3-39
23   Figure 3-13. Inhalation exposures to methanol in human volunteers	3-41
24   Figure 3-14. Blood methanol concentration data from NP and pregnant monkeys	3-44
25   Figure 3-15. Chamber concentration profiles for monkey methanol exposures	3-46
26   Figure 4-1. Hemolymphoreticular neoplasms in male and female Sprague-Dawley rats in
27             formaldehyde	4-110
28   Figure 5-1. Hill model BMD plot of decreased brain weight in male rats at 6 weeks age using
29             modeled AUC of methanol in blood as the dose metric, 1 control mean S.D	5-12
30   Figure 5-2. PODs (in mg/m3) for selected endpoints with corresponding applied UFs	5-15
31   Figure 5-3. All lymphomas versus methanol metabolized (mg/day) for female and male rats. 5-31
32   Figure 5-4. All lymphomas versus Cmax (mg/L)  for female and male rats	5-31
33   Figure 5-5. All lymphomas versus AUC (hr x mg/L) for male and female rats	5-31
34   Figure 5-6. Lympho-immunoblastic Lymphoma minus "lung-only" response in rats of Soffritti et
35             al. (2002a) methanol study versus methanol metabolized (mg/day)	5-42
36   Figure 5-7. Total amount metabolized per day (after periodicity is reached) in a 420 g rat	5-47
37   Figure B-l. Schematic of the PBPK model used to describe the inhalation, oral, and i.v. route
38             pharmacokinetics of MeOH	B-3
39   Figure B-2. Model fits to data sets from GD6, GD7, and GD10 mice for 7-hour inhalation
40             exposures to  1,000-15,000 ppm MeOH	B-7
41   Figure B-3.  Simulation of inhalation exposures to MeOH in NP mice from Perkins et al. (1995a)
42             (8-hour exposures) and Dorman et al. (1995), (6-hour exposures)	B-8
43   Figure B-4.  Oral exposures to MeOH in pregnant mice on GD8 (Dorman et al., 1995) or NP and
44             GDIS	B-10
45   Figure B-5.  Mouse intravenous route MeOH blood kinetics	B-ll
46   Figure B-6.  Mouse model inhalation route sensitivity coefficients for metabolic parameters.B-l5

                                              1-11       DRAFT-DO NOT CITE OR QUOTE

-------
 1   Figure B-7. Mouse model inhalation route sensitivity coefficients for flow rates (QCC: cardiac
 2             output; QPC: alveolar ventilation), and partitioning to the body (PR) compartment are
 3             reported for blood MeOH AUC	B-15
 4   Figure B-8. Mouse model sensitivity coefficients for oral exposures to MeOH	B-16
 5   Figure B-9. Mouse daily drinking water ingestion pattern (A) and resulting predicted blood
 6             concentration for a 6 d/wk exposure (B)	B-17
 7   Figure B-10.  NP rat i.v.-route methanol blood kinetics	B-19
 8   Figure B-l 1. Model fits to data sets from  inhalation exposures to 200 (triangles), 1,200
 9             (diamonds), or 2,000 (squares)  ppm MeOH in male F-344 rats	B-20
10   Figure B-12.  Model fits to data sets from oral exposures to 100 (squares) or 2,500 (diamonds)
11             mg/kg MeOH in female Sprague-Dawley rat	B-22
12   Figure B-13.  Simulated Sprague-Dawley rat inhalation exposures to 500,  1,000, or 2,000 ppm
13             MeOH	B-24
14   Figure B-14.  Simulated rat oral exposures of Sprague-Dawley rats to 65.9, 624.2, or 2,177 mg
15             MeOH/kg/day	B-25
16   Figure B-15. Rat daily drinking water ingestion pattern (A) and resulting predicted blood
17             concentration for a 2-day exposure (B)	B-26
18   Figure B-16. Urinary MeOH elimination  concentration (upper panel) and cumulative amount
19             (lower panel), following inhalation  exposures to MeOH in human volunteers	B-28
20   Figure B-17. Control (nonexposed) blood methanol concentrations	B-30
21   Figure B-18.  Data showing the visual quality of the fit using optimized first-order or Michaelis-
22             Menten kinetics to describe the metabolism of MeOH in humans	B-31
23   Figure B-19.  Inhalation exposures to MeOH in human volunteers	B-33
24   Figure B-20.  Predicted 24-hour AUC (upper left), Cmax (upper right), and amount metabolized
25             (lower) for MeOH inhalation exposures in the mouse (average over a 10-day exposure
26             at 7 hours/day) and humans (steady-state values for a continuous exposure)	B-36
27   Figure B-21.  Fit of the model to i.v. data using different clearance and uptake parameter
28             optimizations	B-41
29   Figure B-22. Fit of the model to oral data using different clearance and uptake parameter
30             optimizations	B-42
31   Figure B-23. Fit of the model to inhalation data using different clearance and uptake parameter
32             optimizations	B-42
33   Figure B-24.  Simulated human oral exposures to 0.1 mg MeOH/kg/-day comparing the first few
34             days for four exposure scenarios	B-92
35   Figure C-l. Hill model, BMR of 1 Control Mean S.D. - Decreased Brain weight in male rats at 6
36             weeks age versus AUC, Fl Generation inhalational study	C-7
37   Figure C-2. Hill model, BMR of 0.05 relative risk - Decreased Brain weight in male rats at 6
38             weeks age versus AUC, FI Generation inhalational study	C-12
39   Figure C-3. Hill model, BMR of 1 Control Mean S.D. - Decreased Brain weight in male rats at 8
40             weeks age versus Cmax, Gestation only  inhalational study	C-18
41   Figure C-4. Exponential model, BMR of 0.05 relative risk - Decreased Brain weight in male
42             rats at 8 weeks age versus Cmax, Gestation only inhalational study	C-23
43   Figure C-5. Nested Logistic Model, 0.1 Extra Risk - Incidence of Cervical Rib in Mice versus
44             Cmax Methanol, GD 6-15  inhalational study	C-32
45   Figure C-6. Nested Logistic Model, 0.05  Extra Risk - Incidence of Cervical  Rib in Mice versus
46             Cmax Methanol, GD 6-15  inhalational study	C-39
47   Figure C-7. 3rd Degree Polynomial Model, BMR of 1 Control Mean S.D. -  VDR in female
48             monkeys using AUC blood  methanol as the dose metric	C-44
                                               1-12       DRAFT-DO NOT CITE OR QUOTE

-------
 1   Figure D-l. Exposure-response data for methanol-induced CR plus SNR malformations in mice
 2             at various concentration-time combinations	D-2
 3   Figure D-2. Internal dose-response relationships for methanol-induced CR plus SNR
 4             malformations in mice at various concentration-time combinations for three dose
 5             metrics	D-3
 6   Figure E-l. Female rat survival	E-20
 7   Figure E-2. Male rats survival	E-20
 8   Figure E-3. Female - Lymphomas lympho-immunoblastic - Multistage Weibull Model -
 9             Approach 1	E-21
10   Figure E-4. Female - All lymphomas - Multistage Weibull Model - Approach 1	E-21
11   Figure E-5. Male - Hepatocellular carcinoma - Multistage Weibull Model - Approach 1	E-22
12   Figure E-6. Male - Lymphomas lympho-immunoblastic - Multistage Weibull Model -
13             Approach 1	E-22
14   Figure E-7. Male - All lymphomas - Multistage Weibull Model - Approach 1	E-23
15   Figure E-8. Female rats -All lymphomas time-to-tumor model fit and Kaplan Meier curves.E-23
16   Figure E-9. Male rats -All lymphomas time-to-tumor model fit and Kaplan Meier curves. ...E-24
17   Figure E-10. Male rats-All lymphomas; dose = amount metabolized (mg/day);  1° multistage
18             model	E-24
19   Figure E-l 1.  Female rat survival	E-25
20   Figure E-12. Male rat survival	E-25
21   Figure E-13. Female rats- pheochromocytomas; dose = amount metabolized (mg/d); 3°
22             multistage model	E-26
23   Figure E-14. Plots for female (top; p= 0.55) and male (bottom; p= 0.21) mice - malignant
24             lymphoma; dose=amount metabolized (mg/d); 2° multistage model	E-26
                                             1-13        DRAFT-DO NOT CITE OR QUOTE

-------
                 LIST OF ACRONYMS
ACGIH      American Conference of Governmental and Industrial Hygienists
ADH        alcohol dehydrogenase
ADH1        alcohol dehydrogenase-1
ADH3        formaldehyde dehydrogenase-3
AIC         Akaike Information Criterion
ALD         aldehyde dehydrogenase
ALDH2      mitochondrial aldehyde dehydrogenase-2
ALT         alanine aminotransferase
ANO VA     analy si s of vari ance
AP          alkaline phosphatase
AST         aspartate aminotransferase
ATP         adenosine triphosphate
ATSDR      Agency for Toxic Substances and Disease Registry
AUC         area under the curve, representing the cumulative product of time
             and concentration for a substance in the blood
P-NAG       N-acetyl-beta-D-glucosaminidase
BAA        butoxyacetic acid
BAL         butoxyacetaldehyde
BMC        benchmark concentration
BMCL       benchmark concentration, 95% lower bound
BMD        benchmark dose(s)
BMDiso      BMD for response one standard deviation from control mean
BMDL       95% lower bound confidence limit on BMD (benchmark dose)
BMDLiso    BMDL for response one standard deviation from control mean
BMDS       benchmark dose software
BMR        benchmark response
BSO         butathione sulfoximine
BUN        blood urea nitrogen
BW, bw      body weight
Ci pool       one carbon pool
Cmax         peak concentration of a substance in the blood during the
             exposure period
C-section     Cesarean section
CA          chromosomal aberrations
CAR         conditioned avoidance response
CASRN      Chemical Abstracts Service Registry Number
CAT         catalase
CERHR      Center for the Evaluation of Risks to Human Reproduction
CH3OH      methanol
CHL         Chinese hamster lung (cells)
                          1-14
DRAFT-DO NOT CITE OR QUOTE

-------
CHR         contact hypersensitivity response
CI           confidence interval
Cls          clearance rate
CNS         central nervous system
CC>2         carbon dioxide
con-A        concanavalin-A
CR          crown-rump length
CSF         Cancer slope factor
CT          computed tomography
CYP450      cytochrome P450
d, 5, A        delta, difference, change
D2           dopamine receptor
DA          dopamine
DIPE        diisopropylether
DMDC       dimethyl dicarbonate
DNA        deoxyribonucleic acid
DNT         developmental neurotoxicity test(ing)
DOPAC      dihydroxyphenyl acetic acid
DPC         days past conception
DTH         delayed-type hypersensitivity
EFSA        European Food Safety Authority
EKG         electrocardiogram
EO          Executive Order
EPA         U.S. Environmental Protection Agency
ERF         European Ramazzini Foundation
EtOH        ethanol
F            fractional bioavailability
FQ           parental generation
FI           first generation
F2           second generation
F344         Fisher 344 rat strain
FAD         folic acid deficient
FAS         folic acid sufficient
FD          formate  dehydrogenase
FP           folate paire
FR          folate reduced
FRACIN     fraction  inhaled
FS           folate sufficient
FSH         follicular stimulating hormone
y-GT         gamma glutamyl transferase
                           1-15
DRAFT-DO NOT CITE OR QUOTE

-------
g            gravity
g, kg, mg, jig gram, kilogram, milligram, microgram
G6PD       glucose-6-phosphate dehydrogenase
GAP43       growth-associated protein (neuronal growth cone)
GD          gestation day
GFR         glomerular filtration rate
GI           gastrointestinal track
GLM        generalized linear model
GLP         good laboratory practice
GSH         glutathione
HAP         hazardous air pollutant
Hb           hemoglobin
HCHO       formaldehyde
HCOO       formate
Hct          hematocrit
HEC         human equivalent concentration
HED         human equivalent dose
HEI         Health Effects Institute
HH          hereditary hemochromatosis
5_HIAA     5-hydroxyindolacetic acid
HMGSH     S-hydroxymethylglutathione
Hp           haptoglobin
HPA         hypothalamus-pituitary-adrenal (axis)
HPLC       high-performance liquid chromatography
HSDB       Hazardous Substances Databank
HSP70       biomarker of cellular stress
5-HT        serotonin
IL           interleukins
i.p.           intraperitoneal
IPCS         International Programme on Chemical Safety
IQ           intelligence quotient
IRIS         Integrated Risk Information System
IUR         inhalation unit risk
i.v.           intravenous
KI           first order rate loss
K1C         first order clearance of methanol from the blood to the bladder for
             urinary elimination
KAI         first order uptake from the intestine
KAS         first order methanol oral absorption rate from stomach
KBL         rate constant for urinary excretion from bladder
KIA         first order uptake from intestine
                           1-16       DRAFT-DO NOT CITE OR QUOTE

-------
KLH        keyhole limpet hemocyanin
KLL         alternate first order rate constant
Km          substrate concentration at half the enzyme maximum
             velocity (Vmax)
Km2         Michaelis-Menten rate constant for low affinity metabolic
             clearance of methanol
KSI         first order transfer between stomach and intestine
L, dL, mL    liter, deciliter, milliliter
LD50         median lethal dose
LDH        lactate dehydrogenase
LH          luteinizing hormone
LLF         (maximum) log likelihood function
LMI         leukocyte migration inhibition (assay)
LOAEL      lowest-observed-adverse-effect level
M, mM, jiM  molar, millimolar, micromolar
MAA        2-methoxyacetic acid
MCH        mean corpuscular hemoglobin
MCHC       mean corpuscular hemoglobin concentration
MCV        mean cell volume
MeOH       methanol
MLE        maximum likelihood estimate
M-M        Michaelis-Menten
MN         micronuclei
MOA        mode of action
4-MP        4-methylpyrazale messenger RNA
MRI         magnetic resonance imaging
MTBE       methyl tertiary butyl ether
MTX        methotrexate
N2O/O2      nitrous oxide
NAD+        nicotinamide adenine dinucleotide
NADH       reduced form of nicotinamide adenine dinucleotide
NET         nitroblue tetrazolium (test)
NCEA       National Center for Environmental Assessment
ND          not determined
NEDO       New Energy Development Organization (of Japan)
NIEHS       National Institute of Environmental Health Sciences
NIOSH      National Institute of Occupational Safety and Health
NK          natural killer
nmol         nanomole
NOAEL      no-observed-adverse-effect level
NOEL       no-observed-effect level
                           1-17
DRAFT-DO NOT CITE OR QUOTE

-------
NP          nonpregnant
NR          not reported
NRC         National Research Council
NS          not specified
NTB         nitroblue tetra zolium
NTP         National Toxicology Program
NZW        New Zealand white (rabbit)
8-OHdG     8-hydroxydeoxyguanosine
OR          osmotic resistance
OSF         oral slope factor
OU          ocular uterque (each eye)
OXA        oxazolone
P, p          probability
PBPK        physiologically based pharmacokinetic
PEG         polyethylene glycol
PFC         plaque-forming cell
PK          pharmacokinetic
PMN        polymorphonuclear leucocytes
PND         postnatal day
POD         point of departure
ppb, ppb     parts per billion, parts per million
PWG        Pathology Working Group of the NTP of NIEHS
Q wave      the initial  deflection of the QRS complex when such deflection is
QCC         cardiac output
QPC         pulmonary (alveolar) ventilation scaling coefficient
QRS         portion of electrocardiogram corresponding to the depolarization
             of ventricular cardiac cells.
R2           square of the correlation coefficient, a measure of the reliability
             of a linear relationship.
RBC         red blood  cell
RfC         reference concentration
RfD         reference dose
RNA        ribonucleic acid
ROS         reactive oxygen species
S9           microsomal fraction from liver
SAP         serum alkaline phosphatase
s.c.          subcutaneous
SCE         sister chromatid exchange
S.D.         standard deviation
S.E.         standard error
SEM         standard error of mean
                           1-18         DRAFT-DO NOT CITE OR QUOTE

-------
SGPT        serum glutamate pyruvate transaminase
SHE         Syrian hamster embryo
SOD         superoxide dismutase
SOP         standard operating procedure(s)
t            time
Ty2, t/2        half-life
T wave       the next deflection in the electrocardiogram after the QRS
             complex; represents ventricular repolarization
TAME       tertiary amyl methyl ether
TAS         total antioxidant status
Tau         taurine
THF         tetrahydrofolate
TLV         threshold limit value
TNFa        tumor necrosis factor-alpha
TNP-LPS    trinitrophenyl-lipopolysaccharide
TRI         Toxic Release Inventory
U83836E    vitamin E derivative
UF(s)        uncertainty factor(s)
             UF associated with interspecies (animal to human) extrapolation
             UF associated with deficiencies in the toxicity database
             UF associated with variation in sensitivity within the human
             population
UFs         UF associated with subchronic to chronic exposure
Vd           volume of distribution
Vmax         maximum enzyme velocity
VmaxC        maximum velocity of the high-affinity/low-capacity pathway
v/v          volume/volume
VDR         visually directed reaching test
VitC         vitamin C
VYS         visceral yolk sac
WBC        white blood cell
WOE        weight of evidence
w/v         weight/volume
%2           chi square
                           1-19
DRAFT-DO NOT CITE OR QUOTE

-------
                                     FOREWORD

       The purpose of this Toxicological Review is to provide scientific support and rationale
for the hazard and dose-response assessment in IRIS pertaining to chronic exposure to methanol.
It is not intended to be a comprehensive treatise on the chemical or toxicological nature of
methanol.
       The intent of Section 6, Major Conclusions in the Characterization of Hazard and Dose
Response, is to present the major conclusions reached in the derivation of the reference dose,
reference concentration and cancer assessment, where applicable, and to characterize the overall
confidence in the quantitative and qualitative aspects of hazard and dose response by addressing
the quality of data and related uncertainties.  The discussion is intended to convey the limitations
of the assessment and to aid and guide the risk assessor in the  ensuing steps of the risk
assessment process.
       For other general information about this assessment or other questions relating to IRIS,
the reader is referred to EPA's IRIS Hotline at (202) 566-1676 (phone), (202) 566-1749 (fax), or
hotline.iris@epa.gov.
                                          1 -20        DRAFT-DO NOT CITE OR QUOTE

-------
                  AUTHORS, CONTRIBUTORS, AND REVIEWERS
                              CHEMICAL MANAGER
Jeffrey Gift, Ph.D.
National Center for Environmental Assessment
U.S Environmental Protection Agency
Research Triangle Park, NC
                                     AUTHORS
Stanley Barone, Ph.D.
National Center for Environmental Assessment
U.S. Environmental Protection Agency
Washington, DC

Allen Davis, MSPH
National Center for Environmental Assessment
U.S. Environmental Protection Agency
Research Triangle Park, NC

Jeffrey Gift, Ph.D.
National Center for Environmental Assessment
U.S. Environmental Protection Agency
Research Triangle Park, NC

Annette lannucci, M.S.
Sciences International (first draft)
2200 Wilson Boulevard
Arlington, VA

Paul Schlosser, Ph.D.
National Center for Environmental Assessment
U.S. Environmental Protection Agency
Washington, DC
                                  CONTRIBUTORS
Bruce Allen, Ph.D.
Bruce Allen Consulting
101 Corbin Hill Circle
Chapel Hill, SC

Hugh Barton, Ph.D.
National Center for Computational Toxicology
U.S. Environmental Protection Agency
Research Triangle Park, NC
                                        1 -21        DRAFT-DO NOT CITE OR QUOTE

-------
J. Michael Davis, Ph.D.
National Center for Environmental Assessment
U.S. Environmental Protection Agency
Research Triangle Park, NC

Robinan Gentry, M.S.
ENVIRON International
650 Poydras Street
New Orleans, LA

Susan Goldhaber, M.S.
Alpha-Gamma Technologies, Inc.
3301 Benson Drive
Raleigh, NC

Mark Greenberg, Ph.D.
Senior Environmental Employee Program
U.S. Environmental Protection Agency
Research Triangle Park, NC

George Holdsworth, Ph.D.
Oak Ridge Institute for Science and Education (second draft)
Badger Road
Oak Ridge, TN

Angela Howard, Ph.D.
National Center for Environmental Assessment
U.S. Environmental Protection Agency
Research Triangle Park, NC

Connie Meacham, M.S.
National Center for Environmental Assessment
U.S. Environmental Protection Agency
Research Triangle Park, NC

Greg Miller
Office of Policy, Economics & Innovation
U.S. Environmental Protection Agency
Washington, DC

Sharon Oxendine, Ph.D.
Office of Policy, Economics & Innovation
U.S. Environmental Protection Agency
Washington, DC

Torka Poet, Ph.D.
Battelle, Pacific Northwest National Laboratories (Appendix B)
902 Battelle Boulevard
Richland, WA
                                        1 -22        DRAFT-DO NOT CITE OR QUOTE

-------
John Rogers, Ph.D.
National Health & Environmental Effects Research laboratory
U.S. Environmental Protection Agency
Research Triangle Park, NC

Reeder Sams, II, Ph.D.
National Center for Environmental Assessment
U.S. Environmental Protection Agency
Research Triangle Park, NC

Roy Smith, Ph.D.
Air Quality Planning & Standards
U.S. Environmental Protection Agency
Research Triangle Park, NC

Frank Stack
Alpha-Gamma Technologies, Inc.
3301 Benson Drive
Raleigh, NC

Justin TeeGuarden, Ph.D.
Battelle, Pacific Northwest National Laboratories
902 Battelle Boulevard
Richland, WA

Chad Thompson, Ph.D., MBA
National Center for Environmental Assessment
U.S. Environmental Protection Agency
Washington, DC

Cynthia VanLandingham, M.S.
ENVIRON International (Appendix E)
650 Poydras Street
New Orleans, LA

Deborah Wales
National Center for Environmental Assessment
U.S. Environmental Protection Agency
Research Triangle Park, NC

Lutz Weber, Ph.D., DABT
Oak Ridge Institute for Science and Education
Badger Road
Oak Ridge, TN
                                        1 -23        DRAFT-DO NOT CITE OR QUOTE

-------
    Errol Zeiger, Ph.D.
    Alpha-Gamma Technologies, Inc.
    3301 Benson Drive
    Raleigh, NC
                                        REVIEWERS
1          This document has been reviewed by EPA scientists, interagency reviewers from other
2   federal agencies and White House offices.

                                INTERNAL EPA REVIEWERS
    Jane Caldwell, Ph.D.
    National Center for Environmental Assessment
    U.S Environmental Protection Agency
    Washington, DC

    Ila Cote, Ph.D., DABT
    National Center for Environmental Assessment
    U.S Environmental Protection Agency
    Washington, DC

    Robert Dewoskin Ph.D., DABT
    National Center for Environmental Assessment
    U.S Environmental Protection Agency
    Research Triangle Park, NC

    Joyce Donahue, Ph.D.
    Office of Water
    U.S. Environmental Protection Agency
    Washington, DC

    Marina Evans, Ph.D.
    National Health and Environmental Effects Research Laboratory
    U.S Environmental Protection Agency
    Research Triangle Park, NC

    Lynn Flowers, Ph.D., DABT
    National Center for Environmental Assessment
    U.S Environmental Protection Agency
    Research Triangle Park, NC

    BrendaFoos, M.S.
    Office of Children's Health Protection and Environmental Education
    U.S Environmental Protection Agency
    Washington, DC

    Jennifer Jinot, Ph.D.
    National Center for Environmental Assessment

                                            1 -24        DRAFT-DO NOT CITE OR QUOTE

-------
U.S Environmental Protection Agency
Washington, DC

Eva McLanahan, Ph.D.
National Center for Environmental Assessment
U.S Environmental Protection Agency
Research Triangle Park, NC

John Vandenberg, Ph.D.
National Center for Environmental Assessment
U.S Environmental Protection Agency
Research Triangle Park, NC

Debra Walsh, M.S.
National Center for Environmental Assessment
U.S Environmental Protection Agency
Research Triangle Park, NC
                                        1 -25       DRAFT-DO NOT CITE OR QUOTE

-------
                                       1. INTRODUCTION

 1          This document presents background information and justification for the Integrated Risk
 2    Information System (IRIS) Summary of the hazard and dose-response assessment of methanol.
 3    IRIS Summaries may include oral reference dose (RfD) and inhalation reference concentration
 4    (RfC) values for chronic and other exposure durations,  and a carcinogenicity assessment.
 5          The RfD and RfC, if derived, provide quantitative information for use in risk assessments
 6    for health effects known or assumed to be produced through a nonlinear (presumed threshold)
 7    mode of action (MOA). The RfD (expressed in units of milligrams per kilogram per day
 8    [mg/kg-day]) is defined as an estimate (with uncertainty spanning perhaps an order of
 9    magnitude) of a daily exposure to the human population (including sensitive subgroups) that is
10    likely to be without an appreciable risk of deleterious effects during a lifetime. The inhalation
11    RfC (expressed in units of milligrams per cubic meter [mg/m3]) is analogous to the oral RfD but
12    provides a continuous inhalation exposure estimate. The inhalation RfC considers toxic effects
13    for both the respiratory system (portal-of-entry) and for effects peripheral to the respiratory
14    system (extrarespiratory or systemic effects).  Reference values are generally derived for chronic
15    exposures (up to a lifetime), but may also be derived for acute (<24 hours), short-term (>24
16    hours up to 30 days), and subchronic (>30 days up to 10% of lifetime) exposure durations, all of
17    which are derived based on an assumption of continuous exposure throughout the duration
18    specified. Unless specified otherwise,  the RfD and RfC are derived for chronic exposure
19    duration.
20          The carcinogenicity assessment provides information on the carcinogenic hazard
21    potential of the substance in question, and quantitative estimates of risk from oral and inhalation
22    exposure may be derived.  The information includes a weight-of-evidence (WOE) judgment of
23    the likelihood that the agent is a human carcinogen and the conditions under which the
24    carcinogenic effects may be expressed. Quantitative risk estimates may be derived from the
25    application of a low-dose extrapolation procedure. If derived, the oral slope factor is  a plausible
26    upper bound on the estimate of risk per mg/kg-day of oral exposure.  Similarly, an inhalation unit
27    risk (IUR) is a plausible upper bound on the estimate of risk per microgram per cubic meter
28    (ug/m3) air breathed.
29          Development of these hazard identification and  dose-response assessments for methanol
30    has followed the general guidelines for risk assessment as set forth by the National Research
31    Council (NRC) (1983). EPA Guidelines and Risk Assessment Forum Technical Panel Reports
32    that may have been used in the development of this assessment include the following: Guidelines
33   for the Health Risk Assessment of Chemical Mixtures (U.S. EPA, 1986a), Guidelines for
34    Mutagenicity Risk Assessment (U.S. EPA, 1986b), Recommendations for and Documentation of
35    Biological Values for Use in Risk Assessment (U.S. EPA, 1988), Guidelines for Developmental
3 6    Toxicity Risk Assessment (U.S. EPA, 1991), Interim Policy for Particle Size and Limit
                                               1 -26        DRAFT-DO NOT CITE OR QUOTE

-------
 1   Concentration Issues in Inhalation Toxicity Studies (U.S. EPA, 1994a), Methods for Derivation
 2   of Inhalation Reference Concentrations and Application of Inhalation Dosimetry (U.S. EPA,
 3   1994b), Use of the Benchmark Dose Approach in Health Risk Assessment (U.S. EPA, 1995),
 4   Guidelines for Reproductive Toxicity Risk Assessment (U.S. EPA, 1996), Guidelines for
 5   Neurotoxicity Risk Assessment (U.S. EPA, 1998), Science Policy Council Handbook: Risk
 6   Characterization (U.S. EPA, 2000a), Benchmark Dose Technical Guidance Document
 1   (U.S. EPA, 2000b), Supplementary Guidance for Conducting Health Risk Assessment of
 8   Chemical Mixtures (U.S. EPA, 2000c), A Review of the Reference Dose and Reference
 9   Concentration Processes (U.S. EPA, 2002a), Guidelines for Carcinogen Risk Assessment
10   (U.S. EPA, 2005a), Supplemental Guidance for Assessing Susceptibility from Early-Life
11   Exposure to Carcinogens (U.S. EPA, 2005b), Science Policy Council Handbook: Peer Review
12   (U.S. EPA, 2006a), and A Framework for Assessing Health Risks of Environmental Exposures to
13   Children (U. S. EPA, 2006b).
14          The literature search strategy employed for this compound was based on the Chemical
15   Abstracts Service Registry Number (CASRN)  and at least one common name. Any pertinent
16   scientific information submitted by the public to the IRIS  Submission Desk was also considered
17   in the development of this document. The relevant literature was reviewed through January,
18   2009.
                                             1 -27        DRAFT-DO NOT CITE OR QUOTE

-------
                         2. CHEMICAL AND PHYSICAL INFORMATION

 1          Methanol is also known as methyl alcohol, wood alcohol; Carbinol; Methylol; colonial
 2    spirit; Columbian spirit; methyl hydroxide; monohydroxymethane; pyroxylic spirit; wood
 3    naphtha; and wood spirit (Chemfmder, 2002). Some relevant physical and chemical properties
 4    are listed below (Hazardous Substances Data Bank [HSDB], 2002, International Programme on
 5    Chemical Safety [IPCS], 1997).

          Table 2-1.  Relevant physical and chemical properties of methanol
                       CASRN:                     67-56-1
                       Empirical formula:            CHsOH
                       Molecular weight:             32.04
                       Vapor pressure:               160 mmHg at 30 °C
                       Vapor Density:                1.11
                       Specific gravity:              0.7866 g/mL (25 °C)
                       Boiling point:                 64.7 °C
                       Melting point:                 -98 °C
                       Water solubility:              Miscible
                       Log octanol-water partition    -0.82 to -0.68
                       coefficient:
                       Conversion factor (in air):      1 ppm =1.31 mg/m3;
                                                    1 mg/m3 = 0.763 ppm

 6          Methanol is a clear, colorless liquid that has an alcoholic odor (IPCS, 1997).  Endogenous
 7   levels of methanol are present in the human body as a result of both metabolism1 and dietary
 8   sources such as fruit, fruit juices, vegetables and alcoholic beverages,2 and can be measured in
 9   exhaled breath and body fluids (Turner et al., 2006; CERHR 2004; IPCS 1997).  Dietary
10   exposure to methanol also occurs through the intake of some food additives. The artificial
11   sweetener aspartame and the beverage yeast inhibitor dimethyl dicarbonate (DMDC) release
12   methanol as they are metabolized (Stegink et al., 1989). In general, aspartame exposure does not
13   contribute significantly to the background body burden of methanol (Butchko, 2002). Oral,
14   dermal, or inhalation exposure to methanol in the environment, consumer products, or workplace
15   also occur.
16          Methanol is a high production volume chemical with many commercial uses and it is a
17   basic building block for numerous chemicals.  Many of its derivatives are used in the
     1 Methanol is generated metabolically through enzymatic pathways such as the methyltransferase system (Fisher
     et al., 2000).
     2 Fruits and vegetables contain methanol.  Further, ripe fruits and vegetables contain natural pectin, which is
     degraded to methanol in the body by bacteria present in the colon (Siragusa et al., 1988). Increased levels of
     methanol in blood and exhaled breath have also been observed after the consumption of ethanol (Fisher et al., 2000).
                                                2-1         DRAFT-DO NOT CITE OR QUOTE

-------
 1   construction, housing or automotive industries, with its use as a transportation fuel accounting
 2   for -20% of its demand (Methanol Institute, 2006).  Consumer products that contain methanol
 3   include varnishes, shellacs, paints, windshield washer fluid, antifreeze, adhesives, de-icers, and
 4   Sterno heaters.  Methanol is used to produce other chemicals and is among the highest
 5   production volume chemicals in the EPA's Toxic Release Inventory (TRI), which reported that
 6   183,000 pounds of methanol was released or disposed of in the United States in 2003, making
 7   methanol among the top 10 chemicals on the list entitled "On-site and Off-site Reported
 8   Disposed of or Otherwise Released (in pounds), for Facilities in All Industries, for All
 9   Chemicals" (U.S. EPA, 2006c).  Natural gas is the primary feedstock for methanol and as such,
10   the major portion of its production cost. Due to the high cost of natural gas in North America,
11   North America methanol production capacity is down from 9 million metric tons at the turn of
12   the century to less than 3 million metric tons at the end of 2005 and is expected to be nonexistent
13   by 2010 (Methanol Institute, 2006). While production has switched to other regions of the
14   world, demand for methanol is growing steadily in almost all end uses. A large reason for the
15   increase in demand is its use in the production of biodiesel, a low-sulfur, high-lubricity fuel
16   source that is gaining favor in nearly every major region of the world.  Starting from less than
17   100,000 metric tons in 2005, global demand for methanol into biodiesel is expected to reach as
18   much as 1.5 million metric tons by the year 2010. Power generation and fuel cells could also be
19   large end users of methanol in the near future (Methanol Institute, 2006).
                                                2-2        DRAFT-DO NOT CITE OR QUOTE

-------
                                      3. TOXICOKINETICS

      3.1. OVERVIEW
 1          As has been noted, methanol occurs naturally in the human body as a product of
 2    metabolism and through intake of fruits, vegetables, and alcoholic beverages (Turner et al., 2006;
 3    CERHR 2004; IPCS 1997). Table 3-1 summarizes background blood methanol levels in healthy
 4    humans which were found to range from 0.25-4.7 mg/L. One study reported a higher
 5    background blood methanol level in females versus males (Batterman and Franzblau, 1997), but
 6    most studies did not evaluate gender differences. Formate, a metabolite of methanol, also occurs
 7    naturally in the human body (IPCS 1997). Table 3-1  outlines background levels of formate in
 8    human blood. In most cases, methanol and formate blood levels were measured in healthy adults
 9    following restriction of methanol-producing foods from the diet.3
10          The absorption, excretion, and metabolism of methanol are well known and have been
11    consistently summarized in reviews such as CERHR (2004), IPCS (1997), U.S. EPA (1996),
12    Kavet and Nauss (1990), HEI (1987), and Tephly and McMartin (1984).  Therefore, the major
13    portion of this toxicokinetics overview is based upon those reviews.
14          Studies conducted in humans and animals demonstrate rapid absorption of methanol by
15    inhalation, oral, and dermal routes of exposure. Table 3-2 outlines increases in human blood
16    methanol levels following various exposure scenarios. Blood levels of methanol following
17    various exposure conditions have also been measured in monkeys, mice, and rats,  and are
18    summarized in Tables 3-3, 3-4, and 3-5, respectively. Once absorbed, methanol pharmacokinetic
19    (PK) data and physiologically based pharmacokinetic (PBPK) model predictions indicate rapid
20    distribution to all organs and tissues according to water content, as an aqueous-soluble alcohol.
21    Tissue:blood concentration ratios for methanol are predicted to be similar through different
22    exposure routes, though the kinetics will vary depending on exposure route and timing (e.g.,
23    bolus oral exposure versus longer-term inhalation). Because smaller species generally have
24    faster respiration rates relative to body weight than larger species, they are predicted to have a
25    higher rate of increase of methanol concentrations in the body when exposed to the same
26    concentration in air.
27          At doses that do not saturate metabolic pathways, a small  percentage of methanol is
28    excreted directly in urine.  Because of the high blood:air partition coefficient for methanol and
29    rapid metabolism in all species studied, the bulk  of clearance therefore occurs by metabolism,
30    though exhalation and urinary clearance will become more significant when doses or exposures
31    are sufficiently high to saturate metabolism.  Metabolic saturation and the corresponding
32    clearance shift have not been observed in humans and nonhuman primates because doses used
      3 In general, background levels among people who are on normal/non-restricted diets will be higher than those
      reported.
                                                3-1         DRAFT-DO NOT CITE OR QUOTE

-------
 1    were limited to the linear range, but the enzymes involved in primate metabolism are also
 2    saturable.
 3          The primary route of methanol elimination in mammals is through a series of oxidation
 4    reactions that form formaldehyde, formate, and carbon dioxide (Figure 3-1).  As noted in
 5    Figure 3-1, methanol is converted to formaldehyde by alcohol dehydrogenase-1 (ADH1) in
 6    primates and by catalase (CAT) and ADH1 in rodents. Although the first step of metabolism
 7    occurs through different pathways in rodents and nonhuman primates, Kavet and Nauss (1990)
 8    report that the reaction proceeds at similar rates (Vmax =30 and 48 mg/h/kg in rats  and nonhuman
 9    primates, respectively).  In addition to enzymatic clearance, methanol can react with hydroxyl
10    radicals to spontaneously yield formaldehyde (Harris et al., 2003). Mannering et al. (1969) also
11    reported a similar rate of methanol metabolism in rats and monkeys, with 10 and 14% of a 1 g/kg
12    dose oxidized in 4 hours, respectively; the rate of oxidation by mice was about twice as fast, 25%
13    in 4 hours. In an HEI study by Pollack and Brouwer (1996), the metabolism of methanol was
14    2 times as fast in mice versus rats, with a Vmax for elimination of 117 and 60.7 mg/h/kg,
15    respectively. Despite the faster elimination rate of methanol in mice versus rats, mice
16    consistently exhibited higher blood methanol levels than rats when inhaling equivalent methanol
17    concentrations (See Tables 3-4 and 3-5).  Possible explanations for the higher methanol
18    accumulation in mice include faster respiration (inhalation rate/body weight) and increased
19    fraction of absorption by the mouse (Perkins et al., 1995a). Because smaller species generally
20    have faster breathing rates than larger species,  humans would be expected to absorb methanol via
21    inhalation more slowly than rats or mice inhaling equivalent concentrations.  If humans eliminate
22    methanol at a comparable rate to rats and mice, then humans would also be expected to
23    accumulate less methanol than those smaller species. However, if humans eliminate methanol
24    more slowly than rats and mice, such that the ratio of absorption to elimination stays the same,
25    then humans would be expected to accumulate methanol to the same internal concentration but to
26    take longer to reach that concentration.
27          In all species, formaldehyde is rapidly converted to formate, with the half-life for
28    formaldehyde being ~1 minute. Formaldehyde is oxidized to formate by two metabolic
29    pathways (Teng et al., 2001). The first pathway (not shown in Figure  3-1) involves conversion
30    of free formaldehyde to formate by the so-called low-affinity pathway (affinity = l/Km =
31    0.002/|jM) mitochondrial aldehyde dehydrogenase-2 (ALDH2).  The second pathway
32    (Figure 3-1) involves a two-enzyme system that converts glutathione-conjugated formaldehyde
33    (S-hydroxymethylglutathione [HMGSH]) to the intermediate S-formylglutathione, which is
34    subsequently metabolized to formate and glutathione (GSH) by S-formylglutathione hydrolase.4
35    The first enzyme in this pathway, formaldehyde dehydrogenase-3 (ADH3), is rate  limiting, and
36    the affinity of HMGSH for ADH3 (affinity = l/Km = 0.15/nM) is about a 100-fold higher than
      4 Other enzymatic pathways for the oxidation of formaldehyde have been identified in other organisms, but this is
      the pathway that is recognized as being present in humans (Caspi et al., 2006; http://metacvc. org)
                                                3-2        DRAFT-DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
13
14
15
that of free formaldehyde for ALDH2. In addition to the requirement of GSH for ADH3 activity,
oxidation by ADH3 is nicotinamide adenine dinucleotide- (NAD+-)dependent.  Under normal
physiological conditions NAD+ levels are about two orders of magnitude higher than NADH,
and intracellular GSH levels (mM range) are often high enough to rapidly scavenge
formaldehyde (Svensson et al., 1999; Meister and Adnerson, 1983); thus, the oxidation of
HMGSH is favorable. In addition, genetic ablation of ADH3 results in increased formaldehyde
toxicity (Deltour et al., 1999). These data indicate that ADH3 is likely to be the predominant
enzyme responsible for formaldehyde oxidation at physiologically relevant concentrations,
whereas ALDHs likely contribute to formaldehyde clearance at higher concentrations (Dicker
and Cederbaum, 1986).
                  Primates

         Alcohol dehydrogenase
               (ADH1)
      Formaldehyde dehydrogenase
               (ADH3)
      S-formylglutathione hydrolase
        Folate dependent pathway
            (see Figure 3-2)
 CH3OH (Methanol)

         I
       HCHO
   (Formaldehyde)
          ( + GSH)
      HMGSH
(hydroxymethyl-GSH)
         4
(^-formyl glutathione)
     |(-GSH)
  HCOO (Formate)
         4
         I
                                                                  Rodents

                                                               CAT and ADH1
                                                         Formaldehyde dehydrogenase
                                                                  (ADH3)
                                                         S-formylglutathione hydrolase
                                                              CAT-peroxide and
                                                           Folate-dependent pathway
                                                               (see Figure 3-2)
                                CO2 (Carbon dioxide)
    Figure 3-1. Methanol metabolism and key metabolic enzymes in primates and rodents.
       Source: IPCS (1997).

       Rodents convert formate to carbon dioxide (CO2) through a folate-dependent enzyme
system and a CAT-peroxide system (Dikalova et al., 2001). Formate can undergo adenosine
triphosphate- (ATP-) dependent addition to tetrahydrofolate (THF), which can carry either one or
two one-carbon groups. Formate can conjugate with THF to form jV10-formyl-THF and its
                                         3-3         DRAFT-DO NOT CITE OR QUOTE

-------
 1   isomer 7V5-formyl-THF, both of which can be converted to TV5, jV70-methenyl-THF and
 2   subsequently to other derivatives that are ultimately incorporated into DNA and proteins via
 3   biosynthetic pathways (Figure 3-2). There is also evidence that formate generates CO2" radicals,
 4   and can be metabolized to CC>2 via CAT and via the oxidation of jV10-formyl-THF (Dikalova
 5   etal.,2001).
                                        Cytoplasm
     Mitochondria
10-formylTHF
ff \C~~^r forarcate^
methenyiTHF THF
\>£- — • senne "Sfi
^(mSHMT
methyleneTHF glycine ~


MTHFDl //
yyfFDH
~ fonrate,_/V (
// \,Q methenyiTHF
THF ' Z ii
..,i~..v. serine — -sS^c-MWT* H
=ss glycine •* methyleneTHF d
I
1 MTHf'R
5-methylTHF
horaocysteme
AdoHyc



	 y dTMP

- — ^
^x
Methionine
)
                                                              AdoMet ^
          Figure 3-2. Folate-dependent formate metabolism. Tetrahydrofolate (THF)-mediated
          one carbon metabolism is required for the synthesis of purines, thymidylate, and
          methionine.
            Source:  Montserrat et al. (2006).

 7          Unlike rodents, formate metabolism in primates occurs solely through a folate-dependent
 8   pathway. Black et al. (1985) reported that hepatic THF levels in monkeys are 60% of that in rats,
 9   and that primates are far less efficient in clearing formate than are rats and dogs.  Studies
10   involving [14C]formate suggest that -80% is exhaled as 14CO2, 2-7% is excreted in the urine, and
11   -10% undergoes metabolic incorporation (Hanzlik et al., 2005, and references therein).  Mice
12   deficient in formyl-TFIF dehydrogenase exhibit no change in LDso (via intraperitoneal [i.p.]) for
13   methanol or in oxidation of high doses of formate.  Thus it has been suggested that rodents
14   efficiently clear formate via folate-dependent pathways, peroxidation by CAT, and by an
15   unknown third pathway; conversely, primates do not appear to exhibit such  capacity and are
16   more sensitive to metabolic acidosis following methanol poisoning (Cook et al., 2001).
17          Blood methanol and formate levels measured in humans under various exposure
18   scenarios are reported in Table 3-2.  As noted in Table 3-2, 75-minute to 6-hour exposures of
19   healthy humans to 200 parts per million (ppm) methanol vapors, the American Council of
                                               3-4
DRAFT-DO NOT CITE OR QUOTE

-------
 1    Governmental Industrial Hygienists (ACGIH) threshold limit value (TLV) for occupational
 2    exposure (ACGIH, 2000), results in increased levels of blood methanol but not formate. A
 3    limited number of monitoring studies indicate that levels of methanol in outdoor air are orders of
 4    magnitude lower than the TLV (IPCS, 1997). Table 3-3 indicates that exposure of monkeys to
 5    600 ppm methanol vapors for 2.5 hours increased blood methanol but not blood formate levels.
 6    Normal dietary exposure to aspartame, which releases 10% methanol during metabolism, is
 7    unlikely to significantly increase blood methanol or formate levels (Butchko et al., 2002).  Data
 8    in Table 3-2 suggest that exposure to high concentrations of aspartame is unlikely to increase
 9    blood formate levels; no increase in blood formate levels were observed in adults ingesting
10    "abusive doses" (100-200 mg/kg) of aspartame (Stegink et al., 1981). Kerns et al. (2002)
11    studied the kinetics of formate in 11 methanol-poisoned patients (mean initial  methanol level of
12    57.2 mmol/L or 1.83 g/L) and determined an elimination half-life of 3.4 hours for formate.
13    Kavet and Nauss (1990) estimated that a methanol dose of 11 mM or 210 mg/kg is needed to
14    saturate folate-dependent metabolic pathways in humans.  There are no data on blood methanol
15    and formate levels following methanol exposure of humans with reduced ADH activity or
16    marginal folate tissue  levels, a possible concern regarding sensitive populations. As discussed in
17    greater detail in Section 3.2, a limited study in folate-deficient monkeys demonstrated no
18    increase in blood formate levels following exposure to 900 ppm methanol vapors for 2 hours.  In
19    conclusion, limited available data suggest that typical occupational, environmental, and dietary
20    exposures are likely to increase baseline blood methanol but not formate levels in most humans.

      3.2. KEY STUDIES
21          Some recent toxicokinetic and metabolism studies (Burbacher et al.,  2004a,  1999b;
22    Medinsky et al., 1997; Pollack and Brouwer, 1996; Dorman et al., 1994) provide key information
23    on interspecies differences,  methanol metabolism during gestation, metabolism in the nonhuman
24    primate, and the impact of folate deficiency on the accumulation of formate.
25          As part of an effort to develop a physiologically based toxicokinetic  model for methanol
26    distribution in pregnancy, Pollack and Brouwer (1996) conducted a large study that compared
27    toxicokinetic differences in pregnant and nonpregnant (NP) rats and mice. Methanol disposition5
28    was studied in Sprague-Dawley rats and CD-I mice that were exposed to 100-2,500 mg/kg of
29    body weight pesticide-grade methanol in saline by intravenous (i.v.) or oral routes.  Exposures
30    were conducted in NP rats and mice, pregnant rats on gestation days (GD)7, GDI4, and GD20,
31    and pregnant mice on  GD9  and GDIS. Disposition was also studied in pregnant rats and mice
32    exposed to 1,000-20,000 ppm methanol vapors  for 8 hours.  Three to five animals were
33    examined at each dose and exposure condition.
      5 Methanol concentrations in whole blood and urine were determined by gas chromatography with flame ionization
      detection (Pollack and Kawagoe, 1991)
                                                3-5         DRAFT-DO NOT CITE OR QUOTE

-------
 1          Based on the fit of various kinetic models to methanol measurements taken from all
 2    routes of exposure, the authors concluded that high exposure conditions resulted in nonlinear
 3    disposition of methanol in mice and rats.6  Both linear and nonlinear pathways were observed
 4    with the relative contribution of each pathway dependent on concentration. At oral doses of
 5    100-500 mg/kg of body weight, methanol was metabolized to formaldehyde and then formic
 6    acid through the saturable nonlinear pathway. A parallel,  linear route characteristic of passive-
 7    diffusion accounted for an increased fraction of total elimination at higher concentrations.
 8    Nearly 90% of methanol elimination occurred through the linear route at the highest oral dose of
 9    2,500 mg/kg of body weight.
10          Oral exposure resulted in rapid and essentially complete absorption of methanol. No
11    significant change in blood area under the curve (AUC) methanol was seen between NP and
12    GD7, GDI4 and GD20 rats exposed to single oral gavage doses of 100 and 2,500 mg/kg, nor
13    between NP and GD9 and GDIS mice at 2,500 mg/kg.  The data as a whole suggested that the
14    distribution of orally and i.v. administered methanol was similar in rats versus mice and in
15    pregnant rodents versus NP rodents with the following exceptions:
16       •  There was a statistically significant increase in the ratio of apparent volume of
17          distribution (Vd) to fractional bioavailability (F) by -20% (while F decreased but not
18          significantly), between NP and GD20 rats exposed to 100 mg/kg orally. However, this
19          trend was not seen in rats or mice exposed to 2,500 mg/kg, and the result in rats at
20          100 mg/kg could well be a statistical artifact since both  Vd and F were being estimated
21          from the same data, making the model effectively  over-parameterized.
22       •  There were statistically significant decreases in the fraction of methanol absorbed by the
23          fast process (resulting in a slower rise to peak blood concentrations, though the peak is
24          unchanged) and Vmax for clearance between NP and GDIS mice. No such differences
25          were  observed between NP and GD9 mice.
26       •  The authors estimated a twofold  higher Vmax for methanol elimination in mice versus rats
27          following oral administration of 2,500 mg/kg methanol, suggesting that similar oral doses
28          would result in lower methanol concentrations in the mouse versus rat.

29          Methanol penetration from maternal blood to the fetal compartment was examined in
30    GD20 rats by microdialysis.7 A plot of the amniotic concentration versus maternal blood
31    concentration (calculated from digitization of Figure 17 of Pollack and Brouwer [1996] report) is
32    shown in Figure 3-3.  The ratio is slightly less than 1:1 (dashed line in plot) and appears to be
33    reduced  with increasing methanol concentrations, possibly due  to decreased blood flow to the
      6 A model incorporating parallel linear and nonlinear routes of methanol clearance was required to fit the data from
      the highest exposure groups.
      7 Microdialysis was conducted by exposing the uterus (midline incision), selecting a single fetus in the middle of the
      uterine horn and inserting a microdialysis probe through a small puncture in the uterine wall proximal to the head of
      the fetus.
                                                3-6         DRAFT-DO NOT CITE OR QUOTE

-------
 1    fetal compartment. Nevertheless, this is a very minor departure from linearity, consistent with a
 2    substrate such as methanol that penetrates cellular membranes readily and distributes throughout
 3    total body water.
                        50001
                     O)
                        4000
                     c
                     O
                     £  3000
                                                                      ,'   *
                                     = -4E-05>? + 1.0782x
                                      1000     2000     3000     4000     5000
                                    Maternal blood concentration (mg/L)	
          Figure 3-3. Plot of fetal (amniotic) versus maternal methanol concentrations in GD20
          rats.   Data extracted from Figure 17 by digitization, and amniotic concentration
          obtains as ("Fetal Amniotic Fluid/Maternal Blood Methanol")x("Maternal
          Methanol").
             Source:  Pollack and Brouwer (1996).

 4           Inhalation exposure resulted in less absorption in both rats and mice as concentrations of
 5    methanol vapors increased, which was hypothesized to be due to decreased breathing rate and
 6    decreased absorption efficiency from the upper respiratory tract.8 Based on blood methanol
 7    concentrations measured following 8-hour inhalation exposures to concentrations ranging from
 8    1,000-20,000 ppm, the study authors (Pollack and Brown,  1996) concluded that methanol
 9    accumulation in the mouse occurred at a two- to threefold greater rate compared to the rat. They
10    speculated that faster respiration rate and more complete absorption in the nasal cavity of mice
11    may explain the higher methanol accumulation and greater sensitivity to developmental toxicity
12    (see Section 4.3.2).
      8 Exposed mice spent some exposure time in an active state, characterized by a higher ventilation rate, and the
      remaining time in an inactive state, with lower (~i/2 of active) ventilation. The inactive ventilation rate was
      unchanged by methanol exposure, but the active ventilation showed a statistically significant methanol-
      concentration-related decline. There was also some decline in the fraction of time spent in the active state, but this
      too was not statistically significant.
                                                 3-7         DRAFT-DO NOT CITE OR QUOTE

-------
 1          The Pollack and Brouwer (1996) study was useful for comparing effects in pregnant and
 2   NP rodents exposed to high doses, but the implication of these results for humans exposed to
 3   ambient levels of methanol is not clear (CERHR, 2004).
 4          Burbacher et al. (2004b, 1999a) examined toxicokinetics mMacaca fascicularis
 5   monkeys prior to and during pregnancy.  The study objectives were to assess the effects of
 6   repeated methanol exposure on disposition kinetics, determine whether repeated methanol
 7   exposures result in formate accumulation, and examine the effects of pregnancy on methanol
 8   disposition and metabolism. Reproductive and developmental toxicity associated with this study
 9   were also examined and are discussed in Sections 4.3.2 and 4.4.2. In a 2-cohort design, 48 adult
10   females (6 animals/dose/group/cohort) were exposed to 0, 200, 600, or 1,800 ppm methanol
11   vapors (99.9% purity) for 2.5 hours/day, 7 days/week for 4 months prior to breeding and during
12   the entire breeding and gestation periods. Six-hour methanol clearance studies were conducted
13   prior to and during pregnancy. Burbacher et al. (2004b, 1999a) reported that:
14       •  At no point during pregnancy was there a significant change in endogenous methanol
15          blood levels, which ranged from 2.2-2.4 mg/L throughout.
16       •  PK studies were performed initially (Study 1), after 90 days of pre-exposure and prior to
17          mating (Study 2), between GD66 and GD72 (Study 3), and again between GD126 and
18          GDI32 (Study 4). These studies were analyzed using classical PK (one-compartment)
19          models.
20       •  Disproportionate mean, dose-normalized, and net blood methanol dose-time profiles in
21          the 600 and 1,800 ppm groups suggested saturation of the metabolism-dependent
22          pathway. Data from the 600 ppm group fit a linear model, while data from the 1,800 ppm
23          group fit a Michaelis-Menten model.
24       •  Methanol clearance rates modestly increased between  Study 1 and Study 2 (90 days prior
25          to mating).  This change was attributed to enzyme induction from the subchronic
26          exposure.
27       •  Blood methanol levels were measured every 2 weeks throughout pregnancy, and while
28          there was measurement-to-measurement variation, there was no significant change or
29          trend over the course of pregnancy. There appears to be an upward trend in elimination
30          half-life and corresponding downward trend in blood methanol clearance between Studies
31          2, 3, and 4.  However, the changes are not statistically  significant and the time-courses
32          for blood methanol concentration (elimination phase) appear fairly similar.
33       •  Significant differences between pre-breeding and gestational blood plasma formate levels
34          were observed but were not dose dependent (Table 3-6).
35       •  Significant differences in serum folate levels in periods prior to and during pregnancy
36          were not dose dependent (Table 3-7).
                                               3-8        DRAFT-DO NOT CITE OR QUOTE

-------
 1          An HEI review committee (Burbacher et al., 1999a) noted that this was a quality study
 2    using a relevant species. Although the study can be used to predict effects in adequately
 3    nourished individuals, the study may not be relevant to persons who are folate deficient.
 4          A series of studies by Medinsky et al.  (1997) and Dorman et al. (1994) examined
 5    metabolism and pharmacokinetics of [14C]methanol and [14C]formate in normal and folate-
 6    deficient cynomolgus, M. fascicularis monkeys that were exposed to environmentally relevant
 7    concentrations of [14C]methanol through an endotracheal tube while anesthetized.  In the first
 8    stage of the study, 4 normal 12-year-old cynomolgus monkeys were each exposed to 10, 45, 200,
 9    and 900 ppm [14C]methanol vapors (>98% purity) for 2 hours.  Each exposure was separated by
10    at least 2 months. After the first stage of the study was completed, monkeys were given a folate-
11    deficient diet supplemented with 1% succinylsulfathiozole (an antibacterial sulfonamide used to
12    inhibit folic acid biosynthesis from intestinal bacteria) for 6-8 weeks in order to obtain folate
13    concentrations of <3 ng/mL serum and <120 ng/mL erythrocytes. Folate  deficiency did not alter
14    hematocrit level, red blood cell count, mean corpuscular volume, or mean corpuscular
15    hemoglobin level. The folate-deficient monkeys were exposed to 900 ppm [14C]methanol for
16    2 hours.  The results of the Medinsky et al. (1997) and Dorman et al. (1994) studies showed:
17       •   Dose-dependent changes in toxicokinetics and metabolism did not occur as indicated by a
18          linear relationship between inhaled [14C]methanol concentration and end-of-exposure
19          blood [14C]methanol level, [14C]methanol AUC and total amounts of exhaled
20          [14C]methanol and [14C]carbon dioxide.
21       •   Methanol concentration had no effect  on elimination half-life (<1  hour) and percent
22          urinary [14C]methanol excretion (<0.01%) at all doses.
23       •   Following exposure to 900 ppm methanol, urinary excretion or exhalation of
24          [14C]methanol did not differ significantly between monkeys in the folate sufficient and
25          deficient state.  There was no significant [14C] formate accumulation at any dose.
26       •   Peak blood [14C]formate levels were significantly higher in folate-deficient monkeys, but
27          did not exceed endogenous blood levels reported by the authors to be between 0.1 and
28          0.2 mmol/L (4.6 to 9.2 mg/L).

29          An FIEI review committee (Medinsky et al., 1997) noted that absolute values in this study
30    cannot be extrapolated to humans because the use of an endotracheal tube in anesthetized
31    animals results in an exposure scenario that is not relevant  to humans. However, the data in this
32    study suggest that a single exposure to an environmentally relevant concentration of methanol is
33    unlikely to result in a hazardous elevation in formate levels, even in individuals with moderate
34    folate deficiency.
                                                3-9        DRAFT-DO NOT CITE OR QUOTE

-------
      3.3. HUMAN VARIABILITY IN METHANOL METABOLISM
 1          The ability to metabolize methanol may vary among individuals as a result of genetic,
 2    age, and environmental factors. Reviews by Agarwal (2001), Burnell et al. (1989), Bosron and
 3    Li (1986), and Pietruszko (1980), discuss genetic polymorphisms for ADH.  Class IADH, the
 4    primary ADH in human liver, is a hetero- or homodimer composed of randomly associated
 5    polypeptide units encoded by three separate gene loci (ADH1 A, ADH1B, and ADH1C).
 6    Polymorphisms have been found to occur at the ADH1B (ADH1B*2, ADH1B*3) and ADH1C
 7    (ADH1C*2) gene loci; however, no human allelic polymorphism has been found in ADH1 A.
 8    The ADH1B*2 phenotype is estimated to occur in -15% of Caucasians of European descent,
 9    85% of Asians, and <5% of African Americans.  Fifteen percent of African Americans have the
10    ADH1B*3 phenotype, while it is found in <5% of Caucasian Europeans and Asians. To date,
11    there are two reports of polymorphisms in ADH3 (Cichoz-Lach et al., 2007; Hedberg et al.,
12    2001), yet the functional consequence(s) for these polymorphisms remains unclear.
13          Although racial and ethnical differences in the frequency of the occurrence of ADH
14    alleles in different populations have been reported, ADH enzyme kinetics (Vmax and Km) have not
15    been reported for methanol.  There is an abundance of information pertaining to the kinetic
16    characteristics of the ADH dimers  to metabolize ethanol in vitro; however, the functional and
17    biological significance is not well understood due to the lack of data documenting metabolism
18    and disposition of methanol or ethanol in individuals of known genotype. While potentially
19    significant, the contribution of ethnic and genetic polymorphisms of ADH to the interindividual
20    variability in methanol disposition and metabolism can not be reliably quantified at this time.
21          Because children generally have higher baseline breathing rates and are more active, they
22    may receive higher methanol doses than adults exposed to equivalent concentrations of any air
23    pollutant (CERHR, 2004). There is evidence that children under 5 years of age have reduced
24    ADH activity. A study by Pikkarainen and Raiha (1967) measured liver ADH activity using
25    ethanol as a substrate and found that 2-month-old fetal livers have -3-4% of adult ADH liver
26    activity.  ADH activity in 4-5 month old fetuses is -10% of adult activity, and an infant's activity
27    is -20% of adult activity.  ADH continues to increase in children with age and reaches a level
28    that is within adult ranges at  5 years of age. Adults were found to have great variation in ADH
29    activity (1,625-6,530/g liver wet weight or 2,030-5,430 mU/100 mg soluble protein).  Smith
30    et al. (1971) also compared liver ADH activity in 56 fetuses (9-22 weeks gestation), 37 infants
31    (premature to <1 year old), and 129 adults (>20 years old) using ethanol as a substrate. ADH
32    activity was 30% of adult activity in fetuses and 50% of adult activity in infants.  There is
33    evidence that some human infants  are able to efficiently eliminate methanol  at high exposure
34    levels, however,  possibly via CAT (Tran et al., 2007).
35          ADH3 exhibits little or no activity toward small alcohols, thus the previous studies
36    provide no information about the ontogeny of formaldehyde clearance. While such data on

                                               3-10         DRAFT-DO NOT  CITE OR QUOTE

-------
 1   ADH3 activity does not exist, ADH3 mRNAis abundantly expressed in the mouse fetus (Ang
 2   et al., 1996) and is detectible in human fetal tissues (third trimester), neonates and children
 3   (Hines et al., 2002; Estonius et al., 1996).
 4          As noted earlier in this section, folate-dependent reactions are important in the
 5   metabolism of formate. Individuals who are commonly folate deficient include those who are
 6   pregnant or lactating, have gastrointestinal (GI) disorders, have nutritionally inadequate diets, are
 7   alcoholics, smoke, have psychiatric disorders, have pernicious anemia, or are taking folic acid
 8   antagonist medications such as some antiepileptic drugs (CERHR, 2004; IPCS, 1997). Groups
 9   which are known to have increased incidence of folate deficiencies include Hispanic and African
10   American women, low-income elderly, and mentally ill elderly (CERHR, 2004). A
11   polymorphism in methylene tetrahydrofolate reductase reduces folate activity and is found in
12   21% of Hispanics in California and 12% of Caucasians in the United States.  Genetic variations
13   in folic acid metabolic enzymes and folate receptor activity are theoretical causes of folate
14   deficiencies.

          Table 3-1.  Background blood methanol and formate levels in humans
Description of human subjects
12 males on restricted diet (no
methanol-containing or methanol-
producing foods) for 12 hr
22 adults on restricted diet (no
methanol-containing or methanol-
producing foods) for 24 hr
3 males who ate a breakfast with no
aspartame-containing cereals and no
juice
5 males who ate a breakfast with no
aspartame-containing cereals and no
juice (second experiment)
Adults who drank no alcohol for 24 hr
12 adults who drank no alcohol for
24 hr
4 adult males who fasted for 8 hr, drank
no alcohol for 24 hr, and took in no
fruits, vegetables, or juices for 18 hr
30 fasted adults
24 fasted infants
Methanol (mg/L)
mean + S.D.
(Range)
0.570 + 0.305
(0.25-1.4)
1.8 + 2.6
(No range data)
1.82+1.21
(0.57-3.57)
1.93+0.93
(0.54-3.15)
1.8 + 0.7
(No range data)
1.7 + 0.9
(0.4-4.7)
No mean data
(1.4-2.6)
<4
(No range data)
<3.5
(No range data)
Formate (mg/L)
mean + S.D.
(Range)
3.8 + 1.1
(2.2-6.6)
11.2 + 9.1
(No range data)
9.08 + 1.26
(7.31-10.57)
8.78 + 1.82
(5.36-10.83)
No data
No data
No data
19.1
(No range data)
No data
Reference
Cook etal. (1991)
Osterloh et al. (1996);
Chuwers et al. (1995)
Lee et al. (1992)
Lee et al. (1992)
Batterman etal. (1998)
Batterman and
Franzblau (1997)
Davoli etal. (1986)
Stegink etal. (1981)
Stegink etal. (1983)
     Source: CERHR (2004).
                                               3-11
DRAFT-DO NOT CITE OR QUOTE

-------
     Table 3-2.  Human blood methanol and formate levels following methanol exposure
Human subjects;
type of sample collected b'c


Adult males and females
administered aspartame; peak
methanol level and range of
formate levels up to 24 hr after
dosing
Infants administered
aspartame; peak exposure
level
Adult males administered
aspartame; range of peak
serum methanol levels in all
subjects

Males; postexposure samples
Males and females;
postexposure serum levels
Males without exercise;
postexposure blood methanol
and plasma formate
Males with exercise;
postexposure blood methanol
and plasma formate
Females; postexposure
samples
Exposure
route


Oral
Oral


Oral

Inhalation

Inhalation

Inhalation

Inhalation
Inhalation
Exposure
duration or
method


1 dose in
juice
1 dose in
beverage

1 dose in
water

75 min

4hr

6hr

6hr
8hr
Methanol
exposure
concentration

0
3.4mg/kgbwa
10 mg/kg bwa
15 mg/kg bwa
20 mg/kg bwa
0
3.4mg/kgbwa
5 mg/kg bwa
10 mg/kg bwa

0
0.6-0.87
mg/kg bwa
0
191 ppm
0
200 ppm

0
200 ppm
0
200 ppm
0
800 ppm
Blood
methanol
mean or
range
(mg/L)

<4
12.7
21.4
25.8
<3.5
3.0
10.2

1.4-2.6
2.4-3.6

0.570
1.881
1.8
6.5

1.82
6.97
1.93
8.13
1.8
30.7
Blood
formate
mean or
range
(mg/L)
19.1
No data
No data
No data
8.4-22.8
No data


No data
3.8
3.6
11.2
14.3

9.08
8.70
8.78
9.52
No data
Reference


Stegink et al.
(1981)
Stegink et al.
(1983)

Davoli et al.
(1986)
Cook et al.
(1991)
Osterloh
et al. (1996)

Lee et al.
(1992)

Batterman
etal.(1998)
aMethanol doses resulting from intake of aspartame.
bUnless otherwise specified, it is assumed that whole blood was used for measurements.
Information about dietary restrictions is included in Table 3-1.

Source: CERHR (2004).
                                           3-12
DRAFT-DO NOT CITE OR QUOTE

-------
    Table 3-3. Monkey blood methanol and formate levels following methanol exposure


Strain-sex

Monkey; Cynomolgus;
female; mean blood
methanol and range of
plasma formate at 30 min
post daily exposure
during premating, mating,
and pregnancy
Monkey; Rhesus male;
postexposure blood level


Exposure
route



Inhalation


Inhalation


Exposure
duration

2.5 hr/day,
7days/wk
during
premating,
mating, and
gestation
(348 days)
6hr

JMethanol
exposure
concentration


o
200 ppm
600 ppm
1 800 ppm


200 ppm
1,200 ppm
2,000 ppm
Blood

methanol
mean
in mg/L

2 4
5
11
35


3.9
37.6
64.4
Blood

formate
mean
in mg/L

8 7
8.7
8.7
10


5.4-13.2
at all doses


Reference



Burbacher et al.
(2004b, 1999a)


Horton et al.
(1992)
Source: CERHR (2004).
    Table 3-4. Mouse blood methanol and formate levels following methanol exposure

Species/strain/sex

Mouse;CD-l;female;
postexposure plasma
methanol and peak
formate level
Mouse;CD-l;female;
postexposure blood
methanol level



Mouse;CD-l;female;

mean postexposure plasma
methanol level


Mouse;CD-l;female;

plasma level 1 hr
postdosing
Mouse;CD-l;female; peak
plasma level

Exposure
route


Inhalation

Inhalation





Inhalation




Gavage
Oral-
Gavage

Exposure
duration


6 hron
GD8
8hr





GD6-GD15




GD6-GD15
GD8

Methanol exposure
concentration

10,000 ppm
10,000 ppm + 4-MP
15,000 ppm
2,500 ppm
5,000 ppm
10,000 ppm
15,000 ppm
0
1,000 ppm
2,000 ppm

5,000 ppm
7,500 ppm
10,000 ppm
15,000 ppm


4,000 mg/kg bw
l,500mg/kgbw
1,500 mg/kg bw +
4-MP
Blood
methanol
mean
(mg/L)
2,080
2,400
7,140
1,883
3,580
6,028
11,165
1.6
97
537

1,650
3,178
4,204
7,330


3,856
1,610
1,450
Blood
formate
mean
(mg/L)
28.5
23
34.5
No data





No data




No data
35
43

Reference


Dorman et al.
(1995)
Pollack and
Brouwer (1996);
Perkins et al.
(1995a)




Rogers et al.
(1993a)




Dorman et al.
(1995)
4-MP=4-methylpyrazole
Source: CERHR (2004).
                                       3-13
DRAFT-DO NOT CITE OR QUOTE

-------
    Table 3-5.  Rat blood methanol and formate levels following methanol exposure
Species;strain/sex:
type of sample
collected

Rat; Sprague-Dawley ;
female; postexposure
blood methanol level
on 3 days

Rat;Sprague-Dawley;
female; postexposure
blood methanol level

Rat;LongEvans;female;
postexposure plasma
levelonGD7-GD12
Rat;LongEvans;female;
1 hr postexposure
blood level

Rat;Long-Evans;male
and female; 1 hr
postexposure blood
level in pups
Rat/Fischer-3 44/male ;
postexposure blood
level
Rat;Long-Evans;male;
post- exposure serum
level


Rat;Long-Evans;male;
peak blood formate






Rat;Long-Evans;male;
peak blood methanol
and formate




Exposure
route

Inhalation



Inhalation


Inhalation

Inhalation


Inhalation

Inhalation
Inhalation



Inhalation






Oral-
gavage




Exposure
duration

7 hr/day for
19 days



8hr


7 hr/day on
GD7- GD
19
6 hr/day on
GD6-
PND21

6 hr/day on
PND1-PND
71

6hr
6hr



6hr






Single dose




Methanol exposure
concentration

5,000 ppm
10,000 ppm
20,000 ppm

1,000 ppm
5,000 ppm
10,000 ppm
15,000 ppm
20,000 ppm
0
15,000 ppm

4,500 ppm


4,500 ppm

200 ppm
1,200 ppm
2,000 ppm
200 ppm
5,000 ppm
10,000 ppm
OFS
OFS
1,200 ppm-FS
1,200 ppm-FR
2,000 ppm-FS
2,000 ppm-FR
3,500mg/kgbw-FS
3,500mg/kgbw-FP
3,500 mg/kg bw-FR
3,000 mg/kg
bw/day-FS
3, 000 mg/kg bw/day
FR

2,000 mg/kg bw/day
FS
2,000 mg/kg bw/day
FR
Blood
methanol
level in mg/L

1,000-2,170
1,840-2,240
5,250-8,650

83
1,047
1,656
2,667
3,916
2.7-1.8
3,826-3,169

555


1,260

3.1
26.6
79.7
7.4
680-873
1,468


No data





4 800
4,800
4,800




Blood
formate
level in
mg/L
No data



No data


No data

No data


No data

5.4-13.2 at
all doses
No data

80
.3
10 1
8.3
46
8.3
83

Baseline
level
382
860
9.2

9 2
538

Reference

Nelson et al.
(1985)

Pollack and
Brouwer
(1996);
Perkins et al.
(1995a)
Stanton et al.
(1995)




(1996)

Horton et al.
(1992)
Cooper et al.
(19921







Lee et al.
(1994)






FS = Folate sufficient; FR = Folate reduced; FP = Folate paire
Source:  CERHR (2004).
                                         3-14
DRAFT-DO NOT CITE OR QUOTE

-------
          Table 3-6. Plasma formate concentrations in monkeys
Exposure group
Control
200 ppm
600 ppm
1800 ppm
Mean plasma formate level (mg/L) during each exposure period
Baseline
8.3
7.4
6.9
6.4
Pre-breeding
7.8
8.3
7.8
8.7
Breeding
10
9.7
9.2
11
Pregnancy
8.3
7.8
8.7
10
     Source: Burbacher et al. (1999a).
          Table 3-7. Serum folate concentrations in monkeys
Exposure group
Control
200 ppm
600 ppm
1800 ppm
Mean serum folate level (jig/L) during each exposure period
Baseline
14.4
11.9
12.5
12.6
Day 70
Pre-pregnancya
14.0
13.2
15.4
14.8
Day 98
Pre-pregnancya
13.4
12.9
13.4
15.3
Day 55
Pregnancy"
16.0
15.5
14.8
15.9
Day 113
Pregnancy"
15.6
13.4
16.4
15.7
     aNumber of days exposed to methanol
     Source: Burbacher et al. (1999a).

     3.4. PHYSIOLOGIC ALLY BASED TOXICOKINETIC MODELS
 1          In accordance with the needs of this human health risk assessment, particularly the
 2   derivation of human health effect benchmarks from studies of the developmental effects of
 3   methanol inhalation exposure in mice (Rogers et al., 1993a) and rats (NEDO, 1987) and
 4   carcinogenic effects of methanol in rats exposed via drinking water (Soffritti et al., 2002a) and
 5   inhalation (NEDO, 1987, 1985/2008b), mouse and rat models were developed to allow for the
 6   estimation of mouse and rat internal dose metrics. A human model was developed to extrapolate
 7   those internal metrics to inhalation and oral exposure concentrations that would result in the
 8   same internal dose in humans (human equivalent concentrations [HECs] and human equivalent
 9   doses [HEDs]). The procedures used for the development, calibration and use of these models
10   are summarized in this section, with further details provided in Appendix B, "Development,
11   Calibration and Application of a Methanol PBPK Model."

     3.4.1. Model Requirements for EPA Purposes
     3.4.1.1. MO A and Selection of a Dose Metric
12          Dose metrics closely associated with one or more key events that lead to the selected
13   critical effect are preferred for dose-response analyses compared to metrics not clearly
14   correlated.  For instance, internal (e.g., blood, target tissue) measures of dose are preferred over
15   external measures of dose (e.g., atmospheric or drinking water concentrations), especially when,
16   as with methanol, blood methanol concentrations increase disproportionally with dose (Rogers
                                               3-15
DRAFT-DO NOT CITE OR QUOTE

-------
 1    et al., 1993a).  This is likely due to the saturable metabolism of methanol in rodents.  In addition,
 2    respiratory and GI absorption may vary between and within species. Mode of action (MOA)
 3    considerations can also influence whether to model the parent compound with or without its
 4    metabolites for selection of the most adequate dose metric.
 5          As discussed in Section 4, developmental effects following methanol exposures have
 6    been noted in both rats and mice, but are not as evident or clear in primate exposure studies
 7    (Burbacher et al., 2004a; Clary, 2003; Rogers et al., 1993a, 1993b; Andrews et al., 1987; Nelson
 8    et al., 1985), and carcinogenic effects have been observed in a drinking water study of Sprague-
 9    Dawley rats (Soffritti et al., 2002a) and an inhalation study of F344 rats (NEDO, 1985/2008b).
10    The report of the New Energy Development Organization (NEDO, 1987) of Japan, which
11    investigated developmental effects of methanol in rats, indicated that there is a potential that
12    developing rat brain weight is reduced following maternal and neonatal exposures. These
13    exposures included both in utero and postnatal exposures. The methanol PBPK models
14    developed for this assessment do not explicitly describe these exposure routes. Mathematical
15    modeling efforts have focused on the estimation of human equivalent external exposures that
16    would lead to internal blood levels of methanol or its metabolites presumed to be associated with
17    developmental effects as reported in rats (NEDO,  1987) and mice (Rogers et al., 1993a), and
18    carcinogenic effects as reported in rats by Soffritti et al. (2002a).
19          In a recent review of the reproductive and developmental toxicity of methanol, a panel of
20    experts concluded that methanol, not formate, is likely to be the proximate teratogen and
21    determined that blood methanol level is a useful biomarker of exposure (CERHR, 2004; Dorman
22    et al., 1995). The CERHR Expert Panel based their assessment of potential methanol toxicity on
23    an assessment of circulating blood levels (CERHR, 2004).  While recent in vitro evidence
24    indicates that formaldehyde is more embryotoxic than methanol and formate (Harris  et al., 2004,
25    2003), the high reactivity of formaldehyde would limit its unbound and unaltered transport as
26    free formaldehyde from maternal to fetal blood (Thrasher and Kilburn, 2001), and the capacity
27    for the metabolism of methanol to formaldehyde is likely lower in the  fetus and neonate versus
28    adults (see discussion in Section 3.3). Thus, even if formaldehyde is ultimately identified as the
29    proximate teratogen, methanol would likely play a prominent role, at least in terms of transport
30    to the target tissue. Further discussions of methanol metabolism, dose metric selection, and
31    MOAissues are covered in Sections 3.3, 4.6, 4.8 and 4.9.2.
32          It has been suggested that the lymphomas observed in Sprague-Dawley rats following
33    methanol  exposure are associated with formaldehyde because formaldehyde and other
34    compounds that metabolize to formaldehyde have been reported to cause lymphomas in Sprague-
35    Dawley rats (Soffritti et al., 2005). Given the reactivity of formaldehyde, models that predict
36    levels of formaldehyde in the blood are difficult to validate. However, production of
37    formaldehyde or formate following exposure to methanol can be estimated by summing the total

                                               3-16        DRAFT-DO NOT CITE  OR QUOTE

-------
 1    amount of methanol cleared by metabolic processes.9 This metric of formaldehyde or formate
 2    dose has limited value since it ignores important processes that may differ between species, such
 3    as clearance of these two metabolites, but it can be roughly be equated to the total amount of
 4    metabolites produced and may be the more relevant dose metric if formaldehyde is found to be
 5    the proximate toxic moiety.  Thus, both blood methanol and total metabolism metrics are
 6    considered to be important components of the PBPK models. Dose metric selection and MOA
 7    issues are discussed further in Sections 3.3, 4.6, 4.8 and 4.9.2.
      3.4.1.2. Criteria for th e Developm en t ofMethan ol PBPK Models
 8          The development of methanol PBPK models that would meet the needs of this
 9    assessment was organized around a set of criteria that reflect: 1) the MOA(s) being considered
10    for methanol; 2) absorption, distribution, metabolism, and elimination characteristics; 3) dose
11    routes necessary for interpreting toxicity studies or estimating HECs; and 4) general parameters
12    needed for the development of predictive PK models.
13          The criteria with a brief justification are provided below:
14       1) Must simulate blood methanol concentrations and total methanol metabolic clearance.
15          Blood methanol is the recommended dose metric for developmental effects, but total
16          metabolic clearance may be a useful metric, particularly for cancer endpoints.
17       2) Must be capable of simulating experimental blood methanol and total metabolic
18          clearance (mg/day) data for the inhalation route of exposure in mice and rats (a) and
19          humans (b), and the oral route in rats (c) and humans (d). These routes are important for
20          determining dose metrics in the most sensitive test species under the conditions of the
21          toxicity study and in the relevant exposure routes in humans.
22       3) The model code should easily allow designation of respiration rates during inhalation
23          exposures. A standard variable in inhalation route risk assessments is ventilation rate.
24          Blood methanol concentrations will depend strongly on ventilation rate, which varies
25          significantly between species.
26       4) Must address the potential for saturable metabolism of methanol.  Saturable metabolism
27          has the potential to bring nonlinearities into the exposure:tissue  dose relationship.
28       5) Model complexity should be consistent with modeling needs and limitations of the
29          available data.  Model  should adequately describe the biological mechanisms that
30          determine the internal dose metrics (blood methanol and metabolic clearance) to assure
31          that it can be reliably used to  predict those metrics in exposure conditions and scenarios
32          where data are lacking. Compartments or processes should not be added that cannot be
33          adequately characterized by the available data.
      9 This assumption is more likely to be appropriate for formaldehyde than formate as formaldehyde is a direct
      metabolite of methanol.
                                                3-17        DRAFT-DO NOT CITE OR QUOTE

-------
 1          Although the rat and mouse models are useful for the evaluation of the dose metrics
 2    associated with methanol's developmental effects and the relevant toxicity studies, including
 3    gestational exposures, no pregnancy-specific PBPK model exists for methanol, and inadequate
 4    data exists for the development and validation of a fetal/gestational/conceptus compartment.
 5    However, EPA determined that nonpregnancy models for the appropriate species and routes of
 6    exposure could prove to be valuable because levels of methanol in NP, pregnant and fetal blood
 7    are expected to be similar following the same oral or inhalation exposure. Pollack and Brouwer
 8    (1996) determined that methanol distribution in rats and mice following repeated oral and i.v.
 9    exposures up to day 20 of gestation is "virtually unaffected by pregnancy, with the possible
10    exception of the immediate  perinatal period." The critical window for methanol induction of
11    cervical rib malformations in CD-I mice has been identified as occurring between GD6 and GD7
12    (Rogers and Mole, 1997;  Rogers et al., 1993b), a developmental period roughly equivalent to
13    week 3 of human development (Chernoff and Rogers, 2004). Methanol blood kinetics measured
14    during and after inhalation exposure in NP and pregnant mice on GD6-GD10 and GD6-GD15
15    (Perkins et al., 1996, 1995a; Dorman et al., 1995; Rogers et al., 1993a) are also similar. Further,
16    the available data indicate that the maternal blood:fetal partition coefficient is approximately 1 at
17    dose levels most relevant to this assessment (Ward et al., 1997; Horton et al., 1992).  The same
18    has been found in rat (Zorzano et al., 1989; Guerri and Sanchis, 1985) and sheep (Brien et al.,
19    1985; Cumming et al.,  1984) studies of ethanol, a structurally related chemical that also
20    penetrates cellular membranes readily and distributes throughout total body water.
21    Consequently, fetal methanol concentrations are expected to be roughly equivalent to that in the
22    mother's blood. Thus,  pharmacokinetics and blood dose metrics for NP mice and humans are
23    expected to provide reasonable approximations of pregnancy levels  and fetal exposure,
24    particularly during early gestation, that improve upon default estimations from external exposure
25    concentrations.

      3.4.2. Methanol PBPK Models
26          As has been discussed, methanol is well absorbed by both inhalation and oral routes and
27    is readily metabolized to formaldehyde, which is rapidly converted to formate in both rodents
28    and humans.  As was discussed in Section 3.1, the enzymes responsible for metabolizing
29    methanol are different in rodents and humans.  Several rat, mouse and human PBPK  models
30    which attempt to account for these species differences have been published (Fisher et al.,  2000;
31    Ward et al., 1997; Perkins et al.,  1995a; Horton et al.,  1992). In addition, a gestational model for
32    a similar water soluble compound, isopropanol, with the potential to be adapted to methanol
33    pharmacokinetics, was of interest (Gentry et al., 2003, 2002; Clewell et al., 2001). Three PK
34    models (Gentry et al., 2002; Bouchard et al., 2001; Ward et al., 1997) were identified as
35    potentially appropriate for use in animal-to-human extrapolation of methanol metabolic rates and
36    blood concentrations.  An additional methanol PBPK model by Fisher et al. (2000) was
                                               3-18        DRAFT-DO NOT CITE  OR QUOTE

-------
 1    considered principally because it had an important feature - pulmonary compartmentalization
 2    (see below for details) - worth adopting in the final model.
      3.4.2.1. Wardetal. (1997)
 3          The PBPK model of Ward et al. (1997) describes inhalation, oral and i.v. routes of
 4    exposure and is parameterized for both NP and pregnant mice and rats (Table 3-8).  The model
 5    has not been parameterized for humans.
 6          Respiratory uptake of methanol is described as a constant infusion into arterial blood at a
 7    rate equal to the minute ventilation times the inhaled concentration and includes a parameter for
 8    respiratory bioavailability, which for methanol is <100%. This simple approach is nonstandard
 9    for volatile compounds but is expected to be appropriate for a compound like methanol, for
10    which there is little clearance from the blood via exhalation.  Oral absorption is described as a
11    biphasic process,  dependent on a rapid and a slow first-order rate constant. This is conceptually
12    similar to the isopropanol model discussed below (Gentry et al., 2002; Clewell et al., 2001),
13    which also employs slow and fast absorption processes but functionally separates them into
14    stomach and duodenal compartments.
15          Methanol  elimination in the Ward et al. (1997) model is primarily via saturable hepatic
16    metabolism. The parameters describing this metabolism come from the literature, primarily
17    previous work by Ward and Pollack (1996) and Pollack et al. (1993). A first-order elimination of
18    methanol  from the kidney compartment includes a lumped metabolic term that accounts for both
19    renal and  pulmonary excretion.
20          The model adequately fits the experimental blood kinetics of methanol in rat and mice
21    and is therefore suitable for simulating blood dosimetry in the relevant test species and routes of
22    exposure (oral and i.v.). The Ward et al. (1997) model meets criteria 1, 2a, 2c, 3, 4, and 5. The
23    most significant limitation is the absence of parameters for the oral and inhalation routes in the
24    human. A modified version of this model that includes human parameters and a standard PBPK
25    lung compartment might be suitable for the purposes of this assessment.
      3.4.2.2. Bouchard et al. (2001)
26          The Bouchard et al. (2001) model is not actually  a PBPK model but is an elaborate
27    classical PK model, since the transfer rates are not determined from blood flows, ventilation,
28    partition coefficients, and the like. The Bouchard et al. (2001) model uses a single compartment
29    for methanol: a central  compartment represented by a volume of distribution where the
30    concentration is assumed to equal that in blood. The model  was  developed for inhalation and i.v.
31    kinetics only. Methanol is primarily eliminated via saturable metabolism. The model adequately
32    simulates blood kinetics in NP rats and humans following inhalation exposure and in NP rats
33    following i.v. exposure; there is no description for oral absorption.  Because methanol distributes
34    with total body water (Ward  et al., 1997; Horton et al., 1992), this simple model  structure is
35    sufficient for predicting blood concentrations of methanol following inhalation and i.v. dosing.
                                               3-19        DRAFT-DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
       The Bouchard et al. (2001) model has the advantage of simplicity, reflecting the
minimum number of compartments necessary for representing blood methanol pharmacokinetics.
Because volume of distribution can be easily  and directly estimated for water-soluble compounds
like methanol or fit directly to experimental kinetics data, concern over the scalability of this
parameter is absent.  The model has been parameterized for a required human exposure route,
inhalation (Table 3-8). The model meets criteria 1, 2b, 3, 4, and 5 described in Section 3.4.1.2.
However, the Bouchard model has specific and significant limitations.  The model has neither
been parameterized for the mouse, a test species of concern (Table 3-8), nor for the oral route in
humans.  As such, the model cannot be used to conduct the necessary interspecies extrapolation.

    Table 3-8. Routes of exposure optimized in models - optimized against blood
    concentration data

Route
i.v.
Inhalation
Oral
Ward et al.
Mouse
P/NP
P/NP
P/NP
Rat
P/NP
-
NP
Human
-
-
-
Bouchard et al.
Mouse
-
-
-
Rat
NP
NP
-
Human
-
NP
-
10
11
12
13
14
15
16
17
18
19
20
21
22
23
P = Pregnant
NP = Nonpregnant
Source: Bouchard et al. (2001); Ward et al. (1997).

3.4.2.3.  Gentry et al. (2003, 2002) and Clewell et al. (2001)
       The model described in these three papers is for isopropanol, not methanol, and therefore
lacks any immediately useful parameterization for the purposes of a methanol risk assessment.
Although the overall model structure, the description of kinetics for both parent compound and
primary metabolite, gestational compartments, lactational transfer, oral and i.v. routes, etc., are
attractive for application to methanol, this model is not ideal. In particular, the model structure is
more elaborate than necessary; because methanol partition coefficients are near 1 for all tissues
except fat, there is no need to individually represent these tissues. Similarly, a fetal compartment
may not be necessary because methanol kinetics in the fetus (conceptus) is expected to parallel
maternal blood  concentrations in the rodent. However, even if a fetal model was considered
necessary, other than the partition coefficient, there are insufficient data to identify conceptus
compartment parameters for methanol. This model would require the most modification and
parameterization to be useful for methanol risk assessment since parameters  would have to be
estimated for all relevant species (at least mouse and humans) and for several routes of exposure.
Therefore the isopropanol model was not considered further.
                                                3-20
                                                      DRAFT-DO NOT CITE OR QUOTE

-------
      3.4.3. Selected Modeling Approach
 1          As discussed earlier regarding model criteria, fetal methanol concentrations can
 2    reasonably be assumed to equal maternal blood concentration. Thus, as discussed in Section
 3    3.4.2.3, methanol pharmacokinetics and blood dose metrics for NP mice and humans are
 4    expected to improve upon default extrapolations from external exposures as estimates of fetal
 5    exposure during early gestation.  The same level of confidence cannot be placed on the whole-
 6    body rate of metabolism, in particular as a surrogate for formaldehyde dose. Because of
 7    formaldehyde's reactivity and the limited fetal metabolic (ADH) activity (see Sections 3.3 and
 8    4.10.1), fetal formaldehyde concentration increases (from methanol) will probably not equal
 9    maternal increases in formaldehyde concentration.  But since there is no model that explicitly
10    describes formaldehyde concentration in the adult, let alone the fetus, the metabolism metric is
11    the closest one can come to predicting fetal formaldehyde dose.  This metric is expected to be a
12    better predictor of formaldehyde  dose than applied methanol dose or even methanol blood levels,
13    which do not account for species differences in conversion of methanol to formaldehyde.
14          Most of the published rodent kinetic models for methanol describe the metabolism of
15    methanol to formaldehyde as a saturable process but differ in the description of metabolism to
16    and excretion of formate (Bouchard et al., 2001; Fisher et al., 2000; Ward et al.,  1997). The
17    model of Ward et al. (1997) used one saturable and one first-order pathway to describe methanol
18    elimination in mice. The saturable pathway described in Ward et al. (1997) can specifically be
19    ascribed to metabolic formation of formaldehyde in the liver, while the renal first-order
20    elimination described in the model represents nonspecific clearance of methanol (e.g.,
21    metabolism, excretion, or exhalation).  The model of Ward et al. (1997) does not describe
22    kinetics of formaldehyde subsequent to its formation and does not include any description of
23    formate.
24          Bouchard et al. (2001) employed a metabolic pathway for conversion of methanol to
25    formaldehyde and a second pathway described as urinary elimination of methanol in rats and
26    humans.  They then explicitly describe two pathways of formaldehyde transformation, one to
27    formate and the other to "other, unobserved formaldehyde byproducts."  Finally, formate
28    removal is described by two pathways, one to urinary elimination and one via metabolism to CO2
29    (which is exhaled).  All of these metabolic and elimination steps are described as first-order
30    processes, but the explicit descriptions of formaldehyde and formate kinetics significantly
31    distinguish the model of Bouchard et al. (2001) from that of Ward et al. (1997), which only
32    describes methanol.
33          There are two other important distinctions between the Ward et al. (1997) and Bouchard
34    et al. (2001) models. The former is currently capable of simulating blood data for all exposure
35    routes in mice but not humans, while the latter is capable of simulating human inhalation route
36    blood pharmacokinetics but not those in mice.  The Ward et al. (1997) model has more

                                                3-21         DRAFT-DO NOT CITE OR QUOTE

-------
 1   compartments than is necessary to adequately represent methanol disposition but has been fit to
 2   PK data in pregnant and NP mice for all routes of exposure (i.v., oral, and inhalation).  The Ward
 3   et al. (1997) model has also been fit to i.v. and oral route PK data in rats.  Based primarily on the
 4   extensive amount of fitting that has already been demonstrated for this model, it was determined
 5   that a modified Ward et al. (1997) model, with the addition of a lung compartment as described
 6   by Fisher et al. (2000), should be used for the purposes of this assessment.  See Appendix B for a
 7   more complete discussion of the selected modeling approach and modeling considerations.
     3.4.3.1. Available PK Data
 8          Although limited human data are available, several studies exist that contain PK and
 9   metabolic data in mice, rats, and nonhuman primates for model parameterization. Table 3-9
10   contains references that were used to verify the model fits as reported in Ward et al. (1997).

          Table 3-9.  Key methanol kinetic studies for model validation
Reference
Battermanetal., 1998
Batterman & Franzblau 1997
Burbacher, 2004a, 2004b
Osterlohetal.,1996;
drawers etal., 1995;
D'Alessandro etal., 1994
Medinsky et al., 1997;
Dormanetal., 1994
Gonzalez-Quevedo et al.,
2002
Hortonetal., 1992
Perkins et al., 1996, 1995a,
1995b
Pollack and Brouwer, 1996;
Pollack etal., 1993
Rogers and Mole, 1997;
Rogers etal., 1993a;
Sedivec et al., 1981
Ward et al., 1997; Ward and
Pollack, 1996
i.v. dose
(mg/kg)






100 (rats
only)

100-2,500


100, 500
(Rat), 2,500
(Mouse)
Inhalation
(ppm)

800 (8 hr)
0-1,800 (2.5 hr, 4
mo)
200 (4 hr)
10-900 (2 hr)

50-2,000 (6 hr)
1,000-20,000(8
hr)
1,000-20,000(8
hr)
1,000-15,000(7
hr, 10 days)
78-231 (8hr)

Oral/dermal/
IP
Dermal




IP: 2 mg/kg-
day, 2 wk


Oral:
100-2,500
mg/kg


Oral:
2,500 mg/kg
Species
Human Male/female

Monkeys
Cynomolgus
Pregnant, NP
Human Male/female
Monkeys
Cynomolgus Folate
deficient
Rat
Rat & Monkey
Rhesus
Mouse and Rat
Rat: Sprague-
Dawley, & Mouse;
CD-I Pregnant, NP
Mouse
Pregnant
Human
GDI 8 Mouse,
GDI 4 & GD20 Rats
Samples
Blood
Blood,
urine,
exhaled
Blood
Blood, urine
Blood,
urine,
exhaled
Blood
Blood,
urine,
exhaled
Blood, urine
Blood
Blood
Urine,
blood
Blood,
conceptus
Digitized
figures"
Figure 1


Figure 1,
Osterloh
etal., 1996


Figure 7



Figures 2, 3,
6,7,8

     adata obtained from the reported figure
                                               3-22
DRAFT-DO NOT CITE OR QUOTE

-------
      3.4.3.2. Model Structure
 1          A model was developed which includes compartments for alveolar air/blood methanol
 2    exchange, liver, fat, bladder (human simulations) and the rest of the body (Figure 3-4). This
 3    model is a revision of the model reported by Ward et al. (1997), reflecting significant
 4    simplifications (removal of compartments for placenta, embryo/fetus, and extraembrionic fluid)
 5    and three elaborations (addition of an intestine lumen compartment to the existing stomach
 6    lumen compartment, use of a saturable rate of absorption from the stomach (but not intestine),
 7    and addition of a bladder compartment which impacts simulations for human urinary excretion),
 8    while maintaining the ability to describe methanol blood kinetics in mice, rats, and humans.  A
 9    fat compartment was included because it is the only tissue with a tissue:blood partitioning
10    coefficient appreciably different than 1, and the liver is included because it is the primary site of
11    metabolism. A bladder compartment was also added for use in simulating human urinary
12    excretion to capture the difference in kinetics between changes in blood methanol concentration
13    and urinary methanol concentration. The model code describes inhalation, oral, and i.v. dose
14    routes, and  data exist (Table 3-9) that were used to fit parameters and evaluate model predictions
15    for all three of those routes in both mice and rats. In humans, inhalation exposure data were
16    available for model calibration and validation but not oral or i.v. data.  However, oral exposures
17    were simulated in humans, assuming a continuous, zero-order ingestion rate, thereby obviating
18    the need for oral uptake parameters.
19          PK data from exposure routes other than inhalation and oral were used to test or further
20    refine the parameters for methanol clearance. Monkey data were evaluated for insight into
21    primate kinetics. Data from Osterloh et al. (1996) and Sedivec et al. (1981) were used to validate
22    the modified Ward et al. (1997) model parameterization for humans. The fact that optimized
23    human parameters were similar to those predicted in monkeys was  important to the validation
24    process (Bourchard et al., 2001)(see section 3.4.7 and Appendix B). Blood levels of methanol
25    have been reported following i.v., oral, and inhalation exposure in rats and mice and inhalation
26    exposure in nonhuman primates and humans.
27          The metabolic clearance of methanol was represented in mice, rats, and humans by
28    specifying separate rate constants for the species-specific enzymes: two saturable processes for
29    mice and rats10 and one for humans. The requirement for two saturable processes in the mouse
30    and rat models may reflect saturation of CAT and ADH1.  Simulated methanol clearance by these
31    metabolic processes is not linked to production of formaldehyde or formate; it is simply cleared
32    from the methanol mass balance leaving the system.  Metabolism of formaldehyde is not
33    explicitly simulated by the model, and this model tracks neither formate nor formaldehyde.
34    Since the metabolic conversion of formaldehyde to formate is rapid (<1 minute) in all species
      10 The need for two saturable metabolic pathways in the mouse model was confirmed through simulation and
      optimization. High exposure (>2,000 ppm methanol) and low exposure (1,000 ppm methanol) blood data could not
      be fit visually, or by more formal optimization, without the second saturable metabolic pathway.
                                                3-23        DRAFT-DO NOT CITE OR QUOTE

-------
 1   (Kavet and Nauss, 1990), the methanol clearance rate may approximate a formate production
 2   rate, though this has not been verified.
 3
Inhalal
Expos
RlnHHar ^ K1
Diauuel ^ —
KBL| (human only)
Urine IV-"*
KLLor
Vm^Y _-
tion 	
ure
^
Venous Blood
_^— —
f
4
^
rf"
^ — '
Alveolar Air
Lung Blood

Body

Fat

Liver
-»E;


*
xha
Arterial Blood
^
                      Vmax2
                      Km2
       Metabolites
                              Kfec
      /KAI
     ——~~*,
Intestine
^KASorVmAS.KMAS
   i*1
    Stomach
   Oral
Exposure
                           Feces
                                   (rat only)
         Figure 3-4. Schematic of the PBPK model used to describe the inhalation, oral, and
         i.v. route pharmacokinetics of methanol.  KAS, first-order oral absorption rate from
         stomach; VmAS and KMAS, Michaelis-Menten rate constants for saturable
         absorption from stomach; KAI, first-order uptake from the intestine; KSI,
         first-order transfer between stomach and intestine; Vmax and Km and Vmax2 and
         Km2, Michaelis-Menten rate constants for high affinity/low capacity and low
         affinity/high capacity metabolic clearance of methanol; KLL, alternate first-
         order rate constant; KBL, rate constant for urinary excretion from bladder.
         Both metabolic pathways were used to describe methanol clearance in the
         mouse, while a single metabolic pathway describes clearance in the human.
 4
 5          The primary purpose of this assessment is for the determination of noncancer and cancer
 6   risk associated with increases in the levels of methanol or its metabolites (e.g., formate,
 7   formaldehyde). Thus, the focus of model development is on obtaining accurate predictions of
 8   increased body burdens over background. The PBPK models do not describe or account for
 9   background levels of methanol, formaldehyde or formate,  and background levels were subtracted
10   from the reported data before use in model fitting or validation (if not already subtracted by
11   study authors), as described below. This approach is not expected to have a significant impact on
12   PBPK model parameter estimates as background levels of methanol and its metabolites are low
                                            3-24
  DRAFT-DO NOT CITE OR QUOTE

-------
 1   relative to exposure levels used in methanol bioassays. Further, while it is possible that
 2   background levels of methanol or its metabolites contribute to background responses for some
 3   adverse effects, the results of dose-response modeling of cancer endpoints using "background
 4   dose" models suggest that this contribution is relatively small (see discussion in Appendix E,
 5   Section E.4).
     3.4.3.3. Model Par am eters
 6          The EPA methanol model uses a consistent set of physiological parameters obtained
 7   predominantly from the open literature (Table 3-10); the Ward et al. (1997) model employed a
 8   number of data-set specific parameters.11  Parameters for blood flow, ventilation, and metabolic
 9   capacity were scaled as a function of body weight raised to the 0.75 power, according to the
10   methods of Ramsey and Andersen (1984).

          Table 3-10.  Parameters used in the mouse, rat and human PBPK models

Body weight (kg)
Mouse
0.03"
Rat
SD | F344
0.2756
Human
70
Source
Measured/estimated
Tissue volume (% body weight)
Liver
Blood arterial
venous
Fat
Lung
Rest of body
5.5
1.23
3.68
7.0
0.73
72.9
3.7
1.85
4.43
7.0
0.50
73.9
2.6
1.98
5.93
21.4
0.8
58.3
Brown etal. 1997
Calculated'
Flows (L/hr/kg0'75)
Alveolar ventillation''
Cardiac output
25.4
25.4
16.4
16.4
16.5
24.0
Perkins et al. 1995a; Brown et al. 1997;
U.S. EPA, 2004
Percentage of cardiac output
Liver
Fat
Rest of body
25.0
5.0
70.0
Biochemical constants8
VmaxC (mg/hr/kg075)
Km (mg/L)
Vmax2C (mg/hr/kg075)
Km2 (mg/L)
K1C (BW025/hr)
19
5.2
3.2
660
NA
25.0
7.0
68

5.0 0
6.3 NA
8.4 22.3
65 100
NA
22.7
5.2
72.1
1s*
, saturable
order
NA 33.1
NA 23.7
NA
NA
0.0373 0.0342
Brown etal. 1997
Calculated

Fitted
     11 Some data sets provided in the Ward et al., (1997) model code were corrected to be consistent with figures in the
     published literature describing the experimental data.
                                               3-25
DRAFT-DO NOT CITE OR QUOTE

-------

KLLC (BW025/hr/
Mouse
NA
Rat
SD | F344
NA
Human
95.7 NA
Source

Oral absorption
VmASC (mg/hr/kg075)
KMASC (mg/kg)
KSI (hr"1)
KAI (hr1)
Kfec (hr'1)
1830
620
2.2
0.33
0
5570
620
7.4
0.051
0.029
377
620
3.17
3.28
0
Mouse and rat fitted (mouse and human
KMASC assumed = rat); other human
values are those for ethanol from Sultatos et
al. (2004), with VmASC set so that for a
70-kg person VmAS/KMAS = the first-
order constant of Sultatos et al.
Partition coefficients
LiverBlood
Fat:Blood
Blood:Air
Body:Blood
Lung:Blood
KBL (hr1), bladder time-
constant J
FRACIN (%), nhalation
fractional availability
1.06
0.083
1350'
0.66
1
1.06
0.083
1350
0.66
1
NA
0.665
0.20
0.583*
0.142
1626
0.805
1.07
0.564 0.612
0.866*
Ward et al., 1997; Fiserova-Bergerova and
Diaz, 1986
Horton et al., 1992; Fiserova-Bergerova and
Diaz, 1986
Rodent: estimated; human: Fiserova-
Bergerova and Diaz, 1986 (human "body"
assumed = muscle)
Fitted (human)
Rodent: fitted; human
Ernstgard et al., 2005
     NA - Not applicable for that species
     "Both sources of mouse data report body weights of approximately 30 g
     *The midpoints of rat weights reported for each study was used and ranged from 0.22 to 0.33 kg
     The volume of the other tissues was subtracted from 91% (whole body minus a bone volume of approximately 9%) to get
     the volume of the remaining tissues
     ''Minute ventilation was measured and reported for much of the data from Perkins et al. (1996) and the average alveolar
     ventilation (estimated as 2/3 minute ventilation) for each exposure concentration was used in the model. When ventilation
     rates were not available, a mouse QPC (Alveolar Ventilation/BW°75) of 25.4 was used (average from Perkins et al., 1995).
     The QPC used to fit the human data was obtained from U.S. EPA (2004). This QPC was somewhat higher than calculated
     from Brown et al. (1997) (-13 L/hr/kg075)
     eVmax, Km, and Vmax2, Km2 represent the two saturable metabolic clearance processes assumed to occur solely in the liver.
     The Vmax used in the model = VmaxC (mg/kg°75-hr) XBW075. K1C is the first-order loss from the blood for human
     simulations  that represents urinary elimination. Allometric scaling for first-order clearance processes was done as previously
     described (Teeguarden et al., 2005);  The Kl used in the model= K1C / BW°25
     fKLLC - alternate human first-order metabolism rate  (used only when VmaxC = Vmax2C = 0)
     ^uman oral simulations used a zero order dose rate equal to the mg/kg-day dose
     *Human liverblood estimated from correlation to (measured) fatblood, based on data from 28 other solvents
     7Rat partition coefficient used for mice as done by Ward et al. (1997)
     JKBL - a first-order rate constant for clearance from the bladder compartment, used to account for the difference between
     blood kinetics and urinary excretion data as  observed in humans
     *For human  exposures, the fractional availability was from Sedivec et al. (1981), corrected for the fact that alveolar
     ventilation is 2/3 of total respiration rate
     3.4.4. Mouse Model Calibration and Sensitivity Analysis

1            The process by which the mouse, rat, and human inhalation and oral models were
2    calibrated is discussed in more detail in Appendix B, "Development, Calibration and Application
3    of a Methanol PBPK Model."  The calibrated mouse inhalation model predicted blood methanol
4    blood concentration time-course agreed well with measured values in adult mice in the critical

                                                    3-26         DRAFT-DO NOT CITE OR QUOTE

-------
 1   inhalation studies of Rogers and Mole (1997) (Figure 3-5), Perkins et al. (1995a, 1995b), and
 2   Rogers et al. (1993a), as well as in NP and early gestation (GD8) mice of Dorman et al. (1995)
 3   (Figure 3-6).  Parameter values used in the calibrated model are given in Table 3-10.
 4          The mouse model was also calibrated for the oral route by fitting all but one of the rate
 5   constants for oral uptake of methanol to the oral-route blood methanol kinetics of Ward et al.
 6   (1997, 1995).  The best model fit to the mouse oral route blood methanol PK data was obtained
 7   using a two-compartment GI tract model, as depicted in Figure 3-4. Because the oral data in rats
 8   led to the conclusion that a saturable rate of uptake from the stomach lumen was necessary (see
 9   section 3.4.5), the same equation was used for uptake in the mouse. But attempts to identify the
10   uptake saturation constant, KMASC, from the mouse data were unsuccessful; therefore KMASC
11   for the mouse was set equal to the value obtained for rats. Adjusting the other mouse oral uptake
12   parameters gave an adequate fit to those data. This calibration allows inhalation to oral dose-
13   route extrapolations in the mouse, which can then be extrapolated to identify human oral route
14   exposures equivalent to mouse inhalation exposures (if equivalent human exposures exist).
                   10,000
                    1,000
                 I
                 o
                 0.1
                 o
                 o
                 m
100
                      10
                                       — 1000 pprn sim
                                       — 2000 ppm sim
                                       — 5000 ppm sim
                                       — 7500 ppm sim
                                       — 10000 pprn sim
                                         15000 ppm sim
                                       i 1000 ppm data (GD6 & 10)
                                       • 2000 ppm data (GD6 & 10)
                                       • 5000 ppm data (GD6 & 10)
                                       * 7500 ppm data (GD6 & 10)
                                       A 10000 pprn data (GD6 & 10)
                                       T 10000 ppm data (GD7)
                                       p 15000 ppm data (GD6 & 10) J
                                                12     15
                                                 Time (hr)
                                                            IB
                                                                  21
                                                                        24
                                                                              27
          Figure 3-5. Model fits to data sets from GD6, GD7, and GD10 mice for 6- to 7-hour
          inhalation exposures to 1,000-15,000 ppm methanol.  Maximum concentrations are
          from Table 2 in Rogers et al. (1993a). The dataset for GD7 mice exposed to
          10,000 ppm is from Rogers and Mole (1997) and personal communication.
          Symbols are concentration + SEM of a minimum of N=4 mice/concentration.
          Default ventilation rates (Table 3-10) were used to simulate these data.

            Source: Rogers and Mole (1997); Rogers et al. (1993a)
                                               3-27
                                      DRAFT-DO NOT CITE OR QUOTE

-------
   8,000-
   7,000-
J  6,000-
o>
E,  5,000-
O  4,000-
v
*  3,000-
o
I  2-ooo^
   1,000-
      0-
                                              + 2,500 ppm NP mouse
                                              A 5,000 ppm IMP mouse
                                              *_t ^r fW V |J- |-r i i     i • • iff v> ^^v
                                              o 10,000 ppm NP mouse
                                              n 6 hr,  lbrOOO ppm GD 8 mouse (Dorman)
                                               15       20
                                                Time (hr)
                                                                   35
          Figure 3-6. Simulation of inhalation exposures to methanol in NP mice from Perkins
          et al. (1995a) (8-hour exposures) and GD8 mice from Dorman et al. (1995)(6-hour
          exposures). Data points are measured blood methanol levels and lines represent
          PBPK model simulations. Digitizlt (Sharlt! Inc., Greensburg, PA) was used to
          digitize data from Figure 2 of Perkins  et al. (1995a) and Figure 2 from Dorman
          et al. (1995). Default ventilation rates (Table 3-10) were used to simulate the
          Dorman data.  The alveolar ventilation rate for each data set from Perkins et al.
          (1995a) was set equal to the measured value reported in that manuscript.  For
          the 2,500, 5,000, and 10,000 ppm exposure groups, the alveolar ventilation rates
          were 29, 24, and 21 (L/hours/kg0'75), respectively. The cardiac output for these
          simulations was set equal to the alveolar ventilation rate.
            Source:  Dorman et al. (1995); Perkins et al. (1995a).

 1          The parameterization of methanol clearance (high-and low-affinity metabolic pathways)
 2   was also verified by  simulation  of datasets describing the pharmacokinetics of methanol
 3   following i.v. administration.  The results of this calibration of the methanol PBPK model are
 4   described in Appendix B and were generally consistent with both the available inhalation and
 5   oral-route data.  Up to 20 hours postexposure, blood methanol kinetics appears similar for NP
 6   and pregnant mice.  However, some data suggests that clearance in GDIS mice is slower than in
 7   NP and earlier in gestation (GD10 and less), particularly beyond 20 hours postexposure (see the
 8   i.v. and oral data of Ward et al. [1997] in Appendix B).
 9          Intravenous-route blood methanol kinetic data in NP mice were only available for a
10   single i.v. dose of 2,500 mg/kg,  but were available for GDIS mice following administration of a
11   broader range of doses:  100, 500, and 2,500 mg/kg. The i.v. maternal PK data in GDIS mice
12   appeared to show an unexpected dose-dependent nonlinearity in initial blood concentrations.
13   Before discussing the nonlinearity, it is first noted that data values used here were obtained from
14   a computational "command file" provided by Ward et al. (1997).  These values appear to be
                                              3-28        DRAFT-DO NOT CITE OR QUOTE

-------
 1    consistent with the plots in their publication but are inconsistent with some of the values in their
 2    Table 6 (Ward et al., 1997).  In particular, the initial maternal blood concentration (i.e., the Cmax)
 3    after the 2,500 mg/kg i.v. is listed as 4,250 mg/L in their command file but as 3,251 mg/L in their
 4    published table. The corresponding data point in their Figure 5 A is distinctly centered above
 5    4,000 mg/L (digitizing yields 4,213 mg/L), and so must be 4,250 rather than 3,251 mg/L.
 6    Therefore the data values listed in the command file were used in the subsequent analysis, rather
 7    than those in the published table.
 8          After i.v. dosing the ratio of the administered doses to the first concentrations measured
 9    by Ward et al. (1997) (5-minute time points) were  0.588 L/kg, 0.585 L/kg, and 0.397 L/kg at
10    doses of 2,500, 500, and 100 mg/kg, respectively.  The discrepancy between the first two values
11    and the third value suggests  either a dose dependence in the Vd or some source of experimental
12    variability.12 It may be that Vd, which is not impacted by any other PBPK parameters and is only
13    determined by the biochemical partitioning properties of methanol, is 1.5-fold lower at 100
14    mg/kg than at the higher concentrations, while the Vd at 500 and 2,500 mg/kg are exactly as
15    predicted by the PBPK model without adjustment. However, it was found that the PBPK model,
16    obtained with measured partition coefficients and otherwise  calibrated to inhalation data, could
17    adequately fit the data at the nominal dose of 100 mg/kg without other parameter adjustment
18    simply by simulating a dose of 200 mg/kg, as  shown in Figure B-5. The fact that the alternate
19    dose (200 mg/kg) differs by  a factor of 2 from the  nominal dose suggests that the data could also
20    be the result of a simple dilution error in dose  preparation. If the first two of the
21    dose/concentration values were not virtually identical (0.588 and 0.585 L/kg), but instead the
22    500 mg/kg value was more intermediate between those for 2,500 and 100 mg/kg, then a regular
23    dose dependence in Vd would seem more likely. However, based on these values, the U.S. EPA
24    has  concluded that the apparent dose dependency is probably the result of a dosing error and
25    therefore, that dose-dependent parameter changes  (e.g., in the partition coefficients) should not
26    be introduced in an attempt to otherwise better fit these data.
27          Further, the nominal  "nonlinearity" between the maternal blood and conceptus shown in
28    Figure 8 of Ward et al. (1997) is the result of those data being plotted on a log-y/linear-x scale.
29    Replotting the data from Tables 5 and 6 (using the value of 4,250 mg/L from the command file as
30    the GDIS maternal Cmax for  the 2,500 mg/kg)  shows the results to be linear, especially in the
31    low-dose region which is of the most concern  (Figure 3-7). Therefore, the current model uses a
32    consistent set of parameters that are not varied by  dose and fit the 2,500 and 500 mg/kg i.v. data
33    adequately, although they  do not fit the 100 mg/kg i.v. data unless, as noted above, a presumed
34    i.v. dose of 200 mg/kg is employed. With that exception, both the single set of parameters used
35    herein and the assumption that maternal blood methanol is a good metric of fetal exposure are
36    well supported by  the data.
      12 It is possible that Ward et al., (1997) were unaware of that discrepancy because they plotted the results for each
      dose in separate figures, and it only becomes obvious when all the data and simulations are plotted together.
                                                3-29        DRAFT-DO NOT CITE OR QUOTE

-------
10000 -
re
1,1000-
Conceptus AUC (
o 8

A
•„•- ' D
{
k
>
) 1000 2000 3000

" \


+ rat
• mouse
-y = x




4000 5000
Maternal AUC (ug/mL-day)
_ 3500 -
•§ 3000
i 2500
3
y 2000
< 1500
(A
a. 1000
8
§ 500
0
(

B
^
/

n/
) 1000 2000 3000

D




> rat
• mouse
-y = x






4000 5000
Maternal AUC (ug/mL-day)
1
2
3
4
    Figure 3-7. Conceptus versus maternal blood AUC values for rats and mice plotted
    (A) on a log-linear scale, as in Figure 8 of Ward et al. (1997), and (B) on a linear-linear
    scale.  In both panels the line y = x is plotted (dashed line) for comparison. Thus the
    apparent "nonlinear" relationship indicated by Ward et al. (1997) is seen to be
    primarily a simple artifact of the choice of axes. However, as evident in panel
    B, there appears to be some nonlinearity at the two highest doses in the mouse
    (results of 2,500 mg/kg i.v. in GDIS mice and 15,000 ppm exposure to GD8
    mice), where distribution from the dam to the conceptus is below  1:1.

       Source: Ward et al. (1997).

       To summarize the mouse model calibration: using the single set of parameters listed for
the mouse in Table 3-10, the PBPK model has been shown to adequately fit or reproduce
methanol PK data from a variety of laboratories and publications, including both NP mice and
pregnant mice up to GD10.  Two saturable metabolic pathways are thus described by the model
                                            3-30
                                                   DRAFT-DO NOT CITE OR QUOTE

-------
 1    and supported by the data. Also, it is thereby demonstrated that a model based on NP mouse
 2    physiology adequately describes (predicts) dosimetry in the pregnant mouse dam through GD10.
 3    Finally, as illustrated in Figure 3-7b, methanol PK in the conceptus and dam of both mice
 4    (including lower doses at GDIS) and rats (GDI4 and GD20) are virtually identical, except for
 5    the very highest doses in mice. Thus the existing model appears to be adequate for predicting
 6    internal methanol doses, including fetal exposures, at bioassay conditions.
 7          An evaluation of the importance of selected parameters on mouse model  estimates of
 8    blood methanol AUC was performed by conducting a sensitivity analysis using the subroutines
 9    within acslXtreme v2.3 (Aegis Technologies, Huntsville, Alabama). The analysis was conducted
10    by measuring the change in model output corresponding to a 1% change in a given model
11    parameter when all other parameters were held fixed. Sensitivity analyses were  conducted for
12    the inhalation and oral routes. The inhalation route analysis was conducted under the exposure
13    conditions of Rogers and Mole (1997) and Rogers et al. (1993a): 7-hour inhalation exposures at
14    the no-observed-effect level (NOEL) concentration of 1,000 ppm. The oral route sensitivity
15    analysis was conducted for an oral dose of 1,000 mg/kg.
16          The parameters with the largest sensitivity coefficients for the inhalation  route at
17    1,000 ppm (absolute values >1) were pulmonary ventilation scaling coefficient (QPC) and
18    maximum velocity of the high-affinity/low-capacity pathway (VmaxC). The sensitivity
19    coefficient for QPC increases during the exposure period as metabolism begins to saturate.
20    Following oral exposure, mouse blood  methanol AUC was sensitive to the rate constants for oral
21    uptake. Blood AUC was most sensitive to the maximum and saturation rate constants for uptake
22    from the stomach (VmASC and KMASC).  The sensitivity coefficient for VmASC decreased
23    during the first hours after exposure from 1 to less than 0.1 at the end of exposure. Blood
24    methanol AUC was also modestly sensitive to first-order uptake from the intestine (KAI), and
25    first-order transfer between stomach and intestine (KSI), the rate constants  for uptake from the
26    intestine and transfer rates between compartments, respectively. For a more complete
27    description of this sensitivity analysis for the mouse methanol PBPK model see Appendix B.

      3.4.5. Rat Model Calibration
28          The rat model was calibrated to fit data from i.v., inhalation, and oral exposures in rats,
29    using data provided in the command file of Ward et al. (1997) and obtained from figures in
30    Horton et al. (1992) using Digitizlt. Holding other parameters constant, the rat PBPK model was
31    initially calibrated against the entire set of i.v.-route blood PK data (Figure  3-7) by fitting
32    Michaelis-Menten constants for one high-affinity/low-capacity and one low-affinity/high-
33    capacity enzyme to both the Ward et al. (1997) data for Sprague-Dawley (SD) rats and the
34    Horton et al. (1992) data for Fischer 344 (F344) rats, assuming that any difference between the
35    two data sets (100 mg/kg data) were from experimental variability and that a single set of
36    parameters could be fit to data for both strains of rat.  However when the resulting parameters
                                               3-31        DRAFT-DO NOT CITE OR QUOTE

-------
 1    were then used to simulate the F344 inhalation uptake data of Horton et al. (with the fractional
 2    absorption for inhalation, FRACIN, adjusted to fit those data), it was found that the clearance
 3    rate predicted after the end of inhalation exposure was much more rapid than shown by the data.
 4    More careful examination of the i.v. data then revealed that there too the clearance for F344 rats
 5    was slower than for SD rats, and that the metabolic parameters obtained from fitting the
 6    combined i.v. data best represented the SD rat  data.  It was concluded that the combined data set
 7    indicated a true strain difference in metabolic parameters. The metabolic parameters for SD rats
 8    were then obtained by fitting only the Ward  et  al. (1997) i.v. data (both doses).
 9          The 100 mg/kg i.v. data of Horton et al. (1992) were combined with their inhalation data
10    and a simultaneous optimization of the metabolic parameters and FRACIN for F344 rats was
11    attempted over that data set. For this data set,  however, the optimization either converged with
12    the metabolic Vmax for the high affinity (low Km) pathway at zero, or with that Km value
13    increasing to be statistically indistinguishable from the high Km value. Therefore the Vmax for
14    the high affinity pathway was allowed to be zero, the Km for that pathway was not estimated,
15    and only a single Vmax and low affinity (high Km) were fit to those data,  with a simultaneous
16    identification of FRACIN.  Since there are no inhalation data for SD rats, this value of FRACIN
17    was assumed to apply for both strains. The  optimized parameters for both strains of rats are
18    given in Table 3-10.
19          When the model was calibrated using the available inhalation and i.v. data for F344 rats
20    (Horton  et al., 1992), a low fractional absorption of 20% was optimized to best fit the data, vs.
21    66.5% for the mouse.  This lower fractional  absorption is consistent with values presented by
22    Perkins et al. (1995), who also found that the fractional absorption of methanol from inhalation
23    studies was lower in rats than in mice.
24
25
26
27
                                               3-32        DRAFT-DO NOT CITE OR QUOTE

-------
  5000
            2500 mg/kg IV in rats
                 16   24   32   40   48
                  Time (hr)
          100 mg/kg IV in rats
                                         cu
                                        O
                                         o>
                                         o
                                         _o
                                         CD
                  Ward '97 data
                  SD rat model fit
                  Morton '92 data
                  F344 rat model fit
           246
               Time (hr)
 Figure 3-8. NP rat i.v. route methanol blood kinetics. Methanol (MeOH) was infused
into: female Sprague-Dawley rats (275 g; solid diamonds and lines) at target
doses of 100 or 2,500 mg/kg (Ward et al., 1997); or male F-344 rats (220 g; open
triangles and dashed line) at target doses of 100 mg/kg (Horton et al., 1992).
Data points represent measured blood concentrations and lines represent PBPK
model simulations.
Source:  Ward et al. (1997); Horton et al. (1992).

                100
                 10
             o
             o
                0,1
                        2000 ppm data
                      * 1200 ppm data
                      A 200 ppm data
                      ^— 2000 ppm, 7-hr simulation
   X	
— 2000 ppm simulation
— 1200 ppm simulation
— 200  ppm simulation
                                                 10
                    15
                                        Time (hr)
 Figure 3-9. Model fits to data sets from inhalation exposures to 200 (triangles), 1,200
 (diamonds), or 2,000 (squares) ppm methanol in male F-344 rats.  The model was
calibrated against all three sets of concentration data, though it converged to
parameter values that only fit the lower two data sets well. Symbols are
concentrations obtained from Horton et al. (1992) using Digitizlt! Lines
represent PBPK model fits. Since the 2000 ppm data peak occurred at 7 hr, a 7-
hr simulated exposure is also shown for comparison.

Source:  Horton et al. (1992).
                                     3-33
    DRAFT-DO NOT CITE OR QUOTE

-------
 1          Finally, oral absorption parameters were optimized to the oral absorption data reported by
 2    Ward et al. (1997), also using the optimization routines in acslXtreme v2.5.0.6 (Aegis
 3    Technologies, Huntsville, Alabama) (Table 3-10: Figure 3-9).  While the two-compartment GI
 4    model (Figure 3-4) allows for both slow and fast absorption modes, it was not possible to fit both
 5    the 100 mg/kg data and the first several hours of the 2500 mg/kg data with that model structure
 6    using linear absorption and inter-compartment transfer rates. In particular the shorter-time data
 7    for 2500 mg/kg indicate a much slower rate of increase in blood levels than the linear-absorption
 8    model (top, thick line in upper panel of Figure 3-10), but the 100 mg/kg data (lower panel of
 9    Figure 3-10) are indeed consistent with a linear model, showing a rapid rise to a fairly narrow
10    peak, then dropping rapidly. As long as linear rate equations were used, the shape of the
11    absorption curve at 2500 mg/kg would mirror that at 100 mg/kg, but the data  show a clear
12    difference.  It was concluded that the rate of absorption must at least partly saturate at the higher
13    dose, and hence that Michaelis-Menten kinetics should be used.
14          Even with the addition of saturable  absorprtion from the stomach, it was also found that
15    the 2500 mg/kg model  simulations over-predicted all of those data (result not shown) and it was
16    hypothesized that fecal elimination might become significant at such a high exposure level, so a
17    term for fecal elimination from the intestine compartment was added. When that fecal rate
18    constant and the saturable absorption from  the stomach compartment were both used, the
19    resulting fit to the data (thin, dashed line in upper panel) was considerably improved with an
20    almost identical (excellent) fit to the 100 mg/kg data (saturable curve can be distinguished from
21    the linear curve just after the peak is reached in the lower panel of Figure 3-4). For the purpose
22    of scaling across individuals, strains, and species, the Km for absorption from the stomach
23    (KMAS) was assumed to scale in proportion to the stomach (lumen) volume;  i.e., with BW1.
24    The Vmax (VmAS) was assumed to scale as BW0'75, with the result that for low doses the
25    effective linear rate constant (VmAS/KMAS) scales as BW"0'25, which is a standard assumption
26    for linear rates.  Since the quantity  on which the rate depends is the total amount in the stomach
27    (mg methanol), the resulting scaling constant for the Km, KMASC, conveniently has units of
28    mg/kg BW; i.e., the standard units for oral dosing.
                                               3-34        DRAFT-DO NOT CITE OR QUOTE

-------
1
2
3
4
5
6
7
            E
            3E
2,700
2,400 -
2,100-
1,800
1,500-
1,200-
  900-
  600-
  300-
    0-
                                              2500 mg/kg linear uptake simulation
                                              2500 mg/kg saturable uptake simulation
                                              2500 mg/kg data
                                              100 mg/kg linear uptake simulation
                                              100 mg/kg saturable uptake simulation
                                              100 mg/kg data
0
                                        12
                                              18
                                          Time (hr)
                                             24
30
36
            o
            E
            ••_••

            O
            Qj
          100 H

           80

           60

           40
            3   20
                                            — 100 mg/kg linear uptake simulation
                                            	 100 mg/kg saturable uptake simulation
                                            n 100 mg/kg data
               012345678
                                         Time (hr)

    Figure 3-10. Model fits to datasets from oral exposures to 100 and 2,500 mg/kg
    methanol in female Sprague-Dawley rats. Symbols are concentration data obtained
    from the command file.  Lines represent PBPK model fits.
       Source:  Ward et al. (1997).

3.4.6. Human Model Calibration
3.4.6.1. Inhalation Route
       The mouse model was scaled to humans by setting either a standard human body weight
(70 kg) or study-specific body weights and using human tissue compartment volumes and blood
flows, and then calibrated to fit the human inhalation exposure data available from the open
literature, which comprised data from four publications  (Ernstgard et al., 2005; Batterman et al.,
1998; Osterloh et al., 1996; Sedivec  et al., 1981).
       Since the human data included time-course data  for urinary elimination, a first-order rate
of loss of methanol from the blood (Kl) was used to provide an estimate of methanol elimination
to the bladder compartment in humans, and the rate of elimination from that compartment then
                                             3-35
                                                     DRAFT-DO NOT CITE OR QUOTE

-------
 1    characterized by a second constant (KBL).  Note that the total amount eliminated by this route
 2    depends only on Kl, while KBL affects the rate at which the material cleared from the blood
 3    then appears in the urine. Inhalation-route urinary methanol kinetic data described by Sedivec
 4    et al. (1981) (Figure 3-11) was used in the model calibration to inform this rate constant.
 5    Without use of the bladder compartment and rate constant, the fit of the model predictions to the
 6    data in Figure 3-11 is quite poor (results not shown), and a statistical test on the improvement of
 7    fit obtained by introducing the additional parameter (KBL) is significant (p < 0.0001).
 8    Conversion between the PBPK-model-predicted rate of urinary excretion (mg/hours) or
 9    cumulative urinary excretion (mg) and the urine methanol concentration data reported by the
10    authors was achieved by assuming 0.5 mL/hours/kg body weight total urinary output (Horton
11    et al., 1992). The resulting values of K1C and KBL, shown in Table 3-10, differ somewhat
12    depending on whether first-order or saturable liver metabolism is used.  These are only calibrated
13    against a small dataset and should be considered an estimate. Urinary elimination is a minor
14    route of methanol clearance with little impact on blood methanol kinetics.
15          Although the high doses used in the mouse studies clearly warrant the use of a second
16    metabolic pathway with a high Km, the human exposure data all represent lower concentrations
17    and may not require or allow for accurate calibration of a second metabolic pathway. Horton
18    et al. (1992) employed two sets of metabolic rate constants to describe human methanol
19    disposition, similar to the description used for rats and mice, but in vitro studies using monkey
20    tissues with nonmethanol substrates were used as justification for this approach.  Although
21    Bouchard et al. (2001) described their metabolism using Michaelis-Menten metabolism, Starr
22    and Festa (2003) reduced that to an effective first-order equation and showed adequate fits.
23    Perkins et al. (1995a) estimated a Km of 320 + 1273 mg/L (mean + S.E.) by fitting a
24    one-compartment model to data from a single estimated oral dose.  In addition to the extremely
25    high standard error, the large standard error for the associated Vmax (93 + 87 mg/kg/hours)
26    indicates that the set of Michaelis-Menten constants was  not uniquely identifiable using this data.
27    Other Michaelis-Menten constants have been used to describe methanol metabolism in various
28    models for primates (Table 3-11).
29
                                               3-36        DRAFT-DO NOT CITE OR QUOTE

-------
           10

            9 H
   -231 ppm saturable sim
   -157 ppm saturable sim
   -78 ppm saturable sim
   • 231 ppm linear sim
   • 157 linear sim
   • 78 ppm linear sim
    231 ppm data
    157 ppm data
         O)
         E,
         o
         is
         -t
         §
         o
         8
Sedivec et al., urine
   concentration
O   78 ppm data
                                                        20
                  24
           2.5 -
                    Sedivec et al.,
                total urinary excretion
                               8        12       16
                                     Time (h)
         20
                                                  24
Figure 3-11. Urinary methanol elimination concentration (upper panel) and
cumulative amount (lower panel) following inhalation exposures to methanol in
human volunteers.  Data points in lower panel represent estimated total urinary
methanol elimination from humans exposed to 78 (diamonds), 157 (triangles), and 231
(circles) ppm methanol for 8 hours, and lines represent PBPK model simulations.
  Source: Sedivec et al. (1981).
                                    3-37
 DRAFT-DO NOT CITE OR QUOTE

-------
          Table 3-11.  Primate Kms reported in the literature
Km (mg/L)
320 +12733
716 + 4893
278
252+1163
33.9
0.66
23.7 + 8.7"'*
Reference
Jacobsenetal., 1988
Nokeretal., 1980
Makar et al., 1968
Eellsetal., 1983
Hortonetal., 1992
Fisher etal., 2000
(This analysis.)
Note
Human: oral poisoning, estimated dose
Cynomolgus Monkey: 2 g/kg dose
Rhesus Monkey: 0.05-1 mg/kg dose
Cynomolgus Monkey: 1 g/kg dose
PBPK model: adapted from rat Km
PBPK model, Cynomolgus Monkey: 10-900 ppm
PBPK model, human: 100-800 ppm
      "The values reported are mean ± S.D.
      bThis Km was optimized while varying Vmax, K1C, and KBL, from all of the at-rest human inhalation data as a part
      of this project. The S.D. given for this analysis is based on the Optimize function of acslXtreme, which assumes all
      data points are discrete and not from sets of data obtained over time; therefore a true S.D. would be higher. The
      final value reported in Table B-l (21 mg/L) was obtained by sequentially rounding and fixing these parameters,
      then re-optimizing the remaining ones.
      Source: Perkins etal. (1995b).
          Table 3-12.  Parameter estimate results obtained using acslXtreme to fit all human
          data using either saturable or first-order metabolism
Parameters
Michaelis-Menten (optimized)
Km
vmaxc
First order
KLLC
Optimized value

23.7
33.1

95.7
S.D.

8.9
10.1

5.4
Correlation coefficient
-0.994


NA

LLF
-24.1


-31.0

      Note. The S.D.s are based on the Optimize function of acslXtreme v2.3, which assumes all data points are discrete
      and not from sets of data obtained over time. Therefore a true S.D. would be a higher value.
 4           To estimate both Michaelis-Menten and first-order rates, all human data under
 5    nonworking conditions (Batterman et al., 1998; Osterloh et al., 1996; Sedivec et al., 1981) were
 6    used (Table 3-12). The metabolic (first-order or saturable) and urinary elimination constants
 7    were numerically fit to the human datasets, while holding the value for FRACIN at 0.8655
 8    (estimated from the results of Sedivec et al.  [1981]) and holding the ventilation rate constant at
 9    16.5 L/hours/kga75 and QPC at 24 L/hours/kg0'75 (values used by EPA [2000d] for modeling the
10    inhalation-route kinetics of vinyl chloride).  Other human-specific physiological parameters were
11    used,as reported in Table 3-10.  Final fitted parameters that have been used in the saturable
                                                  3-38        DRAFT-DO NOT CITE OR QUOTE

-------
1    model are given in Table 3-10.  The resulting fits of two different possible parameterizations
2    (first-order ["linear"] or optimized Km/Vmax) are shown in Figures 3-11 and 3-12.
3
                                    Batterman et al., blood concentration
                                           (800 ppm exposures)
                                                           -Saturable metabolsim
                                                            First-order metabolism
                                                            2 h data
                                                            1 h data
                                                            30 min data
                                                            Osterloh et al.
                                                         (200 ppm inhalation)
                                      O  Data
                                         Saturable metabolsim
                                         First-order metabolism
                                                Time (hr)
4
5
6
7
         Figure 3-12. Data showing the visual quality of the fit using optimized first-order or
         Michaelis-Menten kinetics to describe the metabolism of methanol in humans. Rate
         constants used for each simulation are given in Table 3-12.
            Source:  Batterman et al. (1998: top); Osterloh et al. (1996: bottom).
       Use of a first-order rate has the advantage of resulting in a simpler (one fewer variable)
model, while providing an adequate fit to the data; however, the saturable model clearly fits
some of the data better. To discriminate the goodness-of-fit resulting of the inclusion of an
additional variable necessary to describe saturable metabolism versus using a single first-order
                                           3-39        DRAFT-DO NOT CITE OR QUOTE

-------
 1    rate, a likelihood ratio test was performed.13 The hypothesis that one metabolic description is
 2    better than another is calculated using the likelihood functions evaluated at the maximum
 3    likelihood estimates. Since the parameters are optimized in the model using the maximum log
 4    likelihood function (LLF), the resultant LLF is used for the statistical comparison of the models.
 5    The equation states that two times the log of the likelihood ratio follows a chi square (%2)
 6    distribution with r degrees of freedom:
 7                   -2[log(/l(modell)//l(model2))] =-2[log/l(modell)-log/l(model2)] = %}
 8    The likelihood ratio test states that if the two times the difference between the maximum LLFs of
 9    the two different descriptions of metabolism is greater than the $ distribution then the model fit
10    has been improved (Devore, 1995; Steiner et al., 1990).
11           At greater than a 99.95% confidence level, using two metabolic rate constants (Km and
12    VmaxC) is preferred over using  a single rate constant (Table 3-13). Forcing the model to use the
13    Km calculated by Perkins  et al.  (1995b) would result in model fits indistinguishable from the
14    first-order case (results not shown).  While the correlation coefficients (Table 3-12) indicate that
15    VmaxC, and Km are highly correlated, that is not unexpected, and the S.D.s (Table B-3) indicate
16    that each is reasonably bounded. If the data were indistinguishable from a linear system, Km in
17    particular would not be so bounded from above since the Michaels-Menten model becomes
18    indistinguishable from a linear model as VmaxC and Km tend to infinity. Further, the internal dose
19    candidate points of departure (PODs), for example the BMDLio for the inhalation-induced brain-
20    weight changes from NEDO (1987)  with methanol blood AUC as the metric, is 90.9 mg-hr/L,
21    which corresponds to an average blood concentration  of 3.8 mg/L.  Therefore, the Michaelis-
22    Menten metabolism rate equation  appears to be sufficiently supported by the existing data with
23    values in a concentration range in  which the nonlinearity has an impact.

          Table 3-13. Comparison of LLFs for Michaelis-Menten and first-order metabolism
LLF (logX) for
M-M
-24.1
LLF (log!) for
1st order
-31.0
LLF
1st versus M-Ma
34.1
X2r (99%
confidence)1"
13.8
X2, (99.95%
confidence)1"
12.22
24
      Note. Models were optimized for all human datasets under non working conditions. M-M: Michaelis-Menten
      Obtained using this equation:  - 2[log /l(model l) - log /l(model 2)]
      bsignificance level at r=l degree of freedom.
25           While the use of Michaelis-Menten kinetics might allow predictions across a wide
26    exposure range (into the nonlinear region), extrapolation above 1,000 ppm is not suggested since
27    the highest human exposure data are for 800 ppm. Extrapolation to higher concentrations is
      13 Models are considered to be nested when the model structures are identical except for the addition of complexity,
      such as the added metabolic rate. Under these conditions, the likelihood ratio can be used to compare the relative
      ability of the two models to describe the data, as described in "Reference Guide for Simusolv" (Steiner et al., 1990).
                                                 3-40         DRAFT-DO NOT CITE OR QUOTE

-------
 1   potentially misleading since the nonlinearity in the exposure-internal-dose relationship for
 2   humans is uncertain above this point.  The use of a BMD or internally applied uncertainty factors
 3   (UFs) should place the exposure concentrations well within the linear range of the model.
 4          The data from Ernstgard et al.  (2005) were used to assess the use of the first-order
 5   metabolic rate constant to a dataset collected under conditions of light work.  Historical measures
 6   of QPC (52.6 L/hours/kga75) and QCC (26 L/hours/kg0'75) for individuals exposed under
 7   conditions of 50 watts of work from that laboratory (52.6 L/hours/kg0'75) (Ernstgard, 2005;
 8   Corley et al., 1994; Johanson et al., 1986) were used for the 2-hour exposure period (Figure. 3-
 9   12). Otherwise, there were no changes in the model parameters (no fitting to these data).  The
10   results are remarkably good, given the lack of parameter adjustment to data collected in a
11   different laboratory and using different human subjects than those to which the model was
12   calibrated.
13
14
15
                                        Ernstgard et al. (2005)
    Figure 3-13. Inhalation exposures to methanol in human volunteers.  Data points
    represent measured blood methanol concentrations from humans (4 males and
    4 females) exposed to 100 ppm (open symbols) or 200 ppm (filled symbols) for 2
    hours during light physical activity. Solid lines represent PBPK model
    simulations with no fitting of model parameters. For the first 2 hours, a QPC
    of 52.6 L/hours/kg075 (Johansen et al., 1986), and a QCC of 26 L/hours/kg075
    (Corley et al., 1994) was used by the model.
       Source:  Ernstgard et al. (2005).

3.4.6.2.  Oral Route
       There were no methanol human data available for calibration or validation of the oral
route for the human model. In the absence of methanol data to estimate rate constants for oral
uptake, human oral absorption parameters reported values for ethanol (Sultatos et al., 2004) are
                                               3-41
                                                     DRAFT-DO NOT CITE OR QUOTE

-------
 1    set in the code, except that saturable absorption from the stomach was retained with the KMASC
 2    equal to the mouse value. The maximum rate of absorption form the stomach, VMASC, was
 3    then set such that for a 70-kg person, VMAS/KM (the effective first-order rate constant at low
 4    doses) matched the first-order absorption rate from Sultatos et al. (0.21 hr"1).  Also, while
 5    Sultatos et al. included a rate of metabolism for ethanol in the stomach, the corresponding fecal
 6    elimination rate was set to zero here, effectively assuming 100% absorption of methanol for
 7    humans. However, human oral dosimetry was described as zero-order uptake, in which
 8    continuous infusion at a constant rate into the stomach equal to the daily dose/24 hours was
 9    assumed and human internal doses were computed at steady state.  Since absorption is 100% for
10    the human model, at steady state the net rate of absorption must equal the rate of infusion to the
11    stomach, irrespective of the other parameter values.  (Changes in the absorption constants simply
12    cause the amount of methanol in each GI compartment at steady state to change until the net rate
13    of absorption from the somtach and intestine equals the rate of infusion.) Thus the human
14    absorption constants were set to what is considered a reasonable estimate, given the lack of
15    human oral PK data, but the simulations are conducted in a way that makes the result insensitive
16    to their values; having human values set does allow for simulations of non-constant infusion,
17    should such be desired. Since the AUC was computed for a continuous oral exposure, its value
18    is just 24 hours times the  steady-state blood concentration at a given oral uptake rate.

      3.4.7. Monkey PK Data  and Analysis
19           In order to estimate internal doses (blood AUCs) for the monkey health-effects study of
20    Burbacher et al. (1999b) and further elucidate the potential differences in methanol
21    pharmacokinetics between NP and pregnant individuals (2nd and 3rd trimester), a focused
22    reanalysis of the data of Burbacher et al. (1999a) was performed. Individual blood concentration
23    measurements prior to and following exposure are shown in scatter plots in Appendix B of
24    Burbacher et al. (1999a).  More specifically, the monkeys in the study were exposed for 2.5
25    hours/day, with the methanol concentration raised to approximately the target concentration for
26    the first 2 hours of each exposure and the last 30 minutes providing a chamber "wash-out"
27    period, when the exposure chamber concentration was allowed to drop to 0.   Blood samples
28    were taken and analyzed for methanol concentration at 30 minutes, 1, 2, 3, 4, and 6 hours after
29    removal from the chamber (or 1, 1.5, 2.5, 3.5, 4.5, and 6.5 hours after the end of active
30    exposure). These data were analyzed to compare the PK in NP versus pregnant animals, and
31    fitted with a simple PK model to estimate 24-hour blood AUC values for each exposure level.
32    Dr. Burbacher graciously provided the original data, which were used in this analysis.
33           Two cohorts of monkeys were examined, but the data (plots) did not indicate a systematic
34    difference between the two, so the data from the two cohorts were combined. The data from the
35    scatter plots of Burbacher et al. (1999a) for the NP (pre-pregnancy), first pregnancy (2nd
36    trimester),  and second pregnancy (3rd trimester) studies are compared in Figure 3-13, along with
                                               3-42       DRAFT-DO NOT CITE OR QUOTE

-------
 1    model simulations (explained below).  Since the pregnancy time points were from animals that
 2    had been previously exposed for 87 days plus the duration of pregnancy to that time point, the
 3    pre-exposed NP animals were used for comparison, rather than naive animals, with the
 4    expectation that effects due to changes in enzyme expression (i.e., induction) from the
 5    subchronic exposure would not be a distinguishing factor.  Note that each exposure group
 6    included a pre-exposure baseline or background measurement, also shown.  To aid in
 7    distinguishing the data visually, the NP data are plotted at times 5 minutes prior to the actual
 8    blood draws and the 3rd trimester at 5 minutes after each blood draw.
 9          Overall there appears to be no significant or systematic difference among the NP and
10    pregnant groups. The solid lines are model simulations calibrated to only the 2nd trimester data
11    (details below), but they just as adequately represent average concentrations for the NP and 3rd
12    trimester data.  Likewise, a PK model calibrated to the NP PK data adequately predicted the
13    maternal methanol concentrations in the pregnant monkeys (results not shown). Since any
14    maternal:fetal methanol differences are expected to be similar in experimental animals and
15    humans (with the maternal:fetal ratio being close to one due to methanol's high aqueous
16    solubility and relatively limited metabolism by the  fetus), the predicted levels for the 2nd
17    trimester maternal blood are used in place of measured or predicted fetal concentrations.
18
19
20
21
22
23
24
25
26
27
28
29
30
31
                                               3-43        DRAFT-DO NOT CITE OR QUOTE

-------
                       1800 ppm
                                    64.0 -i
                       -200      -100
                                        0       100      200
                                            Time (min)
                                                               300
                                                                       400
                        600 ppm
                                    24.0-i
                                    18.0 -
                          ±1/1
                                  '
                       -200     -100
                                                         x P re-pregnancy
                                                         + 2ndTrimester
                                                         - SrdTrimester
                                                        	Simulation
                                        0       100      200
                                            Time (min)
                                                               300
                                                                       400
                       200 ppm
                                    10.0-1
                         _x&i/   §
                            *+    m
                       -200      -100
                                        0       100      200
                                            Time (min)
                                                               300
                                                                       400
1
2
3
4
5
6
    Figure 3-14. Blood methanol concentration data from NP and pregnant monkeys. NP
    and 3rd trimester data are plotted, respectively, at 5 minutes before and after
    actual collection times to facilitate comparison.  Solid line is from simple PK
    model, fit to 2nd trimester data only.
       Source: Burbacher et al. (1999a; Figure B-4).

3.4.7.1. PK Model Analysis for Monkeys
       To analyze and integrate the PK data of Burbacher et al. (1999a), the one-compartment
model for Michaelis-Menten kinetics used by Burbacher et al. (1999a, 1999b) was extended by
the addition of a chamber compartment to capture the kinetics of concentration change in the
exposure chamber, as shown in Figure 3-14. The data in Figure 3-14 (digitized from Figure 5 of
Burbacher et al.,  1999a, 1999b)  show an exponential rise to and fall from the approximate target
concentration during the exposure period.  The use of a single-compartment model for the
                                             3-44
                                                     DRAFT-DO NOT CITE OR QUOTE

-------
 1    chamber allows this dynamic to be captured, so that the full concentration-time course is used in
 2    simulating the monkey internal concentration rather than an approximate step function (i.e. rather
 3    than assuming an instantaneous rise and fall). The pair of equations representing the time-course
 4    in the chamber and monkey are as follows (bolded parameters are fit to data):

 5                        Chamber: dCch/dt = [(CCM- S - Cch> Fch - Rmh]/Vch

 6                   Monkey:  dCmk/dt = [Rmh - Vmax- Cmk/(Km + (?„*)]/(Vmk- BW)

 7                    with Rmh = Cch-Rc-(1000-BW)0-74 -F and Cnet = Cmk + Cbg.

 8       d: delta, change
 9       Ccn: instantaneous chamber concentration (mg/L)
10       t: time (hour)
11       CCM: chamber in-flow methanol concentration (mg/L), which was set to the concentrations
12          corresponding to those reported in Table 2 of Burbacher et al. (1999a), using the
13          "Breeding" column for the NP (87 days pre-exposed; values in  Table 3-14)
14       S:  exposure switch, set to 1 when exposure is on (first 2 hours) and 0 when off
15       Fch:chamber air-flow, 25,200 L/hours, as specified by Burbacher et al. (1999a,1999b)
16       Rinh: net rate of methanol inhalation by the monkeys (mg/hr)
17       VCh (1,220 L): chamber volume, initially set to 1,380 L ("accessible volume" stated by
18          Burbacher et al. [1999a, 1999b]), but allowed to vary below that value to account for
19          volume taken by equipment, monkey, and to  allow for imperfect mixing
20       C^: instantaneous inhalation-induced monkey blood methanol concentration (mg/L); this is
21          added to the measured background/endogenous concentration before comparison to data
22       Vmax (39.3 mg/hr): fitted (nonscaled) Michaelis-Menten maximum elimination rate
23       Km (14.6 mg/L): fitted (nonscaled) Michaelis-Menten saturation constant
24       Vmk (0.75 L/kg):  fitted volume of distribution for monkey
25       BW: monkey body weight (kg); for NP monkeys set to group average values in data of
26          Burbacher et al. (1999a, 1999b; personal communication)
27       RC: allometric scaling factor for total monkey respiration (0.12 L/hours/g0'74 =
28          2 mL/minute/ga74), as used by Burbacher et al. (1999a, 1999b)(note that scaling is to BW
29          in g, not kg)
30       F:  fractional absorption of inhaled methanol, set to 0.6 (60%), the (rounded) value measured
31          in humans by Sedivec etal. (1981); F and Vmk cannot be uniquely identified, given the
32          model structure, so F was set to the (approximate) human value to obtain a realistic
3 3          estimate of Vmk
34       Cnet: net blood concentration, equal to sum of the inhalation-induced concentration (Cmk) and
35          the background blood level (Cbg) (mg/L)
36       Cbg: background (endogenous) methanol concentration, set to the pre-exposure group-
37          specific mean from the data of Burbacher et al. (1999a, 1999b;  personal communication)
                                               3-45       DRAFT-DO NOT CITE OR QUOTE

-------


b
Q.
Q.
C
O
1
0
O
C
O
O




2000 -
1800
1600

1400
1200

1000
800
600

400
onn


c

A
f 1800 ppm


Chamber concentration profile
k
I
600 ppm
-•- -m--m--m---m — m--m--m--m---m — •--,

[,•
' 200 ppm

I I I
J 0.5 1 1.5 2
Time (hr)










A
'•vSv
™"


          Figure 3-15.  Chamber concentration profiles for monkey methanol exposures.  Lines
            are model simulations. Indicated concentrations are target concentrations;
            measured concentrations differed slightly (see Table 3-14).
            Source: Burbacher et al. (1999a).

 1          The model was specifically fit to the 2nd trimester monkey data, assuming that the
 2    parameters were the same for all the exposure groups and concentrations. While the discussion
 3    above and data show little difference between the NP and two pregnancy groups, the 2nd
 4    trimester group was presumed to be most representative of the average internal dosimetry over
 5    the entire pregnancy.  Further, the results of Mooney and Miller (2001) show that developmental
 6    effects on the monkey brain stem following ethanol exposure are essentially identical for
 7    monkeys exposed only during early pregnancy versus full-term, indicating that early pregnancy
 8    is a primary window of vulnerability.
 9          Model simulation results are the lines shown in Figures  3-13 and 3-14.  The model
10    provides a good fit to the monkey blood and chamber air concentration data.  While the chamber
11    volume was treated as a fitted parameter, which was not done by Burbacher et al. (1999a), the
12    chamber concentration data support this estimate. The model does an adequate job of fitting the
13    data for all exposure groups without group-specific parameters. In particular, the data for all
14    exposure levels can be adequately fit using a single value for the volume of distribution (Vmk) as
15    well as each of the metabolic parameters.  While one may be able to show statistically distinct
16    parameters for different groups or exposure levels (by fitting the model separately to each),  as
17    was done by Burbacher et al. (1999a), it is unlikely that such differences are biologically
18    significant, given the fairly large number of data points and the large variability evident in the
19    blood concentration data. Thus, the single set of parameters listed with the parameter
20    descriptions above will be used to estimate internal blood concentrations for the dose-response
                                               3-46
DRAFT-DO NOT CITE OR QUOTE

-------
 1    analysis.  The chamber concentrations for "pregnancy" exposures recorded by Burbacher et al.
 2    (1999a; Table 2) and average body weights for each exposure group at the 2nd trimester time
 3    point were used along with the model to calculate 24-hour blood methanol AUCs (Table 3-14).

          Table 3-14.  Monkey group exposure characteristics
Exposure concentration (ppm)a
206
610
1,822
Group average BW (kg)b
3.46
4.08
3.83
24-hr blood methanol AUC
(mg-hr/L)c
6.73
28.28
138.11
      aFrom Burbacher et al. (1999a,1999b), Table 2, "pregnancy" exposure.
      bFrom Burbacher, original data (personal communication).
      Calculated using the two-compartment PK model as described above.

      3.4.8. Summary and Conclusions
 4          Mouse, rat, and human versions of a methanol PBPK model have been developed and
 5    calibrated to data available in the open literature.  The model simplifies the structure used by
 6    Ward et al. (1997), while adding specific refinements such as a standard lung compartment
 7    employed by Fisher et al. (2000) and a two-compartment GI tract.
 8          Although the endpoints of concern are developmental effects which occur during in utero
 9    and (to a lesser extent) lactational exposure, no pregnancy-specific PBPK model exists for
10    methanol and inadequate data exists for the development and validation of a
11    fetal/gestational/conceptus compartment.  The fact that the unique physiology of pregnancy and
12    the fetus/conceptus are not represented in a methanol model would be important if methanol
13    pharmacokinetics differed significantly during pregnancy or if the observed partitioning of
14    methanol into the fetus/conceptus versus the mother showed a concentration ratio significantly
15    greater than or less than 1. Methanol pharmacokinetics during GD6-GD10 in the mouse are not
16    different from NP mice (Pollack and Brouwer, 1996), and the maternal blood:fetus/conceptus
17    partition coefficient is reported to be near 1 (Ward et al., 1997; Horton et al., 1992). At GDIS in
18    the mouse, maternal blood levels are only modestly different from those in NP animals (see
19    Figures B-4 and B-5 [Appendix B] for examples), and in general the PBPK model simulations
20    for the NP animal match the pregnancy data as well as the nonpregnancy data. Likewise,
21    maternal blood kinetics in monkeys differs little from those in NP animals (see Section 3.4.7 for
22    details).  Further, in both mice and monkeys, to the extent that late-pregnancy blood levels differ
23    from NP for a given exposure, they are higher; i.e., the difference between model predictions and
24    actual concentrations is in the same direction.  These data support the assumption that the ratio of
25    actual target-tissue methanol concentration to (predicted) NP maternal blood concentrations will
26    be about the same across species, and hence, that  using NP maternal blood levels in place of fetal
27    concentrations will not lead to a systematic error when extrapolating risks.

                                               3-47        DRAFT-DO NOT CITE OR QUOTE

-------
 1           The findings in the mouse (similar blood methanol kinetics between NP and pregnant
 2    animals prior to GDIS and a maternal blood:fetal partition coefficient close to 1) are assumed to
 3    be applicable to the rat. However, the critical gestational window for the reduced brain weight
 4    effect observed in the NEDO (1987) rat study is broader than for the mouse cervical rib effect.
 5    In addition, NEDO (1987) rats were exposed not only to methanol gestationally but also
 6    lactationally and via inhalation after parturition. The additional routes of exposure presented to
 7    the pups in this study present uncertainties and suggest that average blood levels in pups might
 8    be greater than those of the dam.
 9           Methanol is transported directly from the maternal circulation to fetal circulation via the
10    placenta, but transfer via lactation involves distribution to the breast tissue, then milk, then
11    uptake  from the pup's GI tract.  Therefore blood or target-tissue levels in the breast-feeding
12    infant or pup are likely to differ more from  maternal levels than do fetal levels. In addition, the
13    health-effects data indicate that most of the effects of concern are due to fetal exposure, with a
14    relatively small influence due to postbirth exposures. Further, it would be extremely difficult to
15    distinguish the contribution of postbirth exposure from pre birth exposure to a given effect in a
16    way that would allow the risk to be  estimated from estimates of both exposure levels, even if one
17    had a lactation/child PBPK model that allowed for prediction of blood (or target-tissue) levels in
18    the offspring. Finally, one would still expect the target-tissue concentrations in the offspring to
19    be closely related to maternal blood levels (which depend on ambient exposure and determine
20    the amount delivered through breast milk),  with the relationship between maternal levels and
21    those in the offspring being similar across species.  Further,  as discussed to a greater extent in
22    Sections 5.1.2 and 5.3, it is likely that the difference in blood levels between rat pups and dams
23    would be similar to the difference between  mothers and human offspring. Therefore, it is
24    assumed that the potential differences between pup and dam blood methanol levels do not have a
25    significant impact on this risk assessment and the estimation of HECs. The use of the full
26    intrahuman UF of 10 is also expected to account for PK differences between children and adults.
27           Therefore, the development  of a lactation/child PBPK model appears not to be necessary,
28    given the minimal change that is likely to result in risk extrapolations, and use of (NP) maternal
29    blood levels as a measure of risk in the offspring is considered preferable over use of default
30    extrapolation methods. In particular, the existing human data allow for predictions of maternal
31    blood levels, which depend strongly on the rate of maternal methanol clearance. Since bottle-fed
32    infants do not receive methanol from their mothers, they are expected to have lower or,  at most,
33    similar overall exposures for a given ambient concentration than the breast-fed infant, so that use
34    of maternal blood levels for risk estimation should also be adequately protective for that group.
35           The model fits to the mouse oral-route methanol kinetic data, using a consistent set of
36    parameters (Figure B-4 in Appendix B), are fairly good for doses of 1,500 mg/kg but
37    underpredict blood levels by 30% or more after a dose of 2,500 mg/kg. In particular, the oral
38    mouse  model consistently underpredicts the amount of blood methanol reported in two studies
                                                3-48        DRAFT-DO NOT CITE OR QUOTE

-------
 1   (Ward et al., 1997, 1995). Ward et al. (1997) utilized a different Vmax for each oral absorption
 2   dataset; the GDIS and the GD8 data from Dorman et al.  (1995) were both fit using a Vmax of
 3   -80 mg/kg/hours (body weights were not listed; the model assumed that GD8 and GDIS mice
 4   were both 30 g; Ward et al. [1997] did not scale by body weight). Additionally, lower partition
 5   coefficients for placenta (1.63 versus 3.28) and embryonic fluid (0.0037 versus 0.77) were used
 6   for GD8 and GDIS. The current refined model adequately fits the oral PK data using a single set
 7   of parameters that is not varied by dose or source of data.
 8          The rat models were able to adequately predict the limited inhalation, oral and i.v.
 9   datasets available. Low-dose exposures were emphasized in model optimization due to their
10   greater relevance to risk assessment. Based on a rat inhalation exposure to 500 ppm, the HEC
11   would be 281 ppm (by applying an AUC of 201.3 [Figure B-12] to Equation 1 of Appendix B).
12          The final mouse, rat, and human methanol PBPK models fit multiple datasets for
13   inhalation, oral, and i.v., from multiple research groups using consistent parameters that are
14   representative of each species but are not varied within species or by dose or source of data.
15   Also, a simple PK model calibrated to NP monkey data,  which were shown to be essentially
16   indistinguishable from pregnant monkey PK data, was used to estimate blood methanol AUC
17   values (internal doses) in that species.  In Section 5, the models and these results are used to
18   estimate chronic human exposure concentrations from internal dose metrics.
                                              3-49        DRAFT-DO NOT CITE OR QUOTE

-------
                                  4. HAZARD IDENTIFICATION

      4.1. STUDIES IN HUMANS - CASE REPORTS, OCCUPATIONAL AND
      CONTROLLED STUDIES
      4.1.1. Case Reports
 1          An extensive library of case reports has documented the consequences of acute
 2    accidental/intentional methanol poisoning. Nearly all have involved ingestion, but a few have
 3    involved percutaneous and/or inhalation exposure. As many of the case reports demonstrate, the
 4    association of Parkinson-like symptoms with methanol poisoning is related to the observation
 5    that lesions in the putamen are a common feature both in Parkinson's disease and methanol
 6    overexposure.  These lesions are commonly identified using computed tomography (CT) or by
 7    Magnetic Resonance Imaging (MRI).  Other areas of the brain (e.g., the cerebrum, cerebellum,
 8    and corpus callosum) also have been shown to be adversely affected by methanol overexposure.
 9    Various therapeutic procedures (e.g., ethanol infusion, sodium bicarbonate or folic acid
10    administration, and hemodialysis) have been used in many of these methanol overexposures,  and
11    the reader is referred to the specific case reports for details in this regard. The reader also is
12    referred to Kraut and Kurtz (2008) and Barceloux et al. (2002) for a more in-depth discussion of
13    the treatments in relation to clinical features of methanol toxicity.  A brief discussion of the
14    terms cited in case report literature follows.
15          Basal ganglia, a group of interconnected subcortical nuclei in each cerebral hemisphere,
16    refers to various structures in the grey  matter of the brain that are intimately involved, for
17    example,  in coordinating motor function, maintaining ocular and respiratory function, and
18    consciousness.  The connectivity within the basal ganglia involves both excitatory and inhibitory
19    neurotransmitters such as dopamine (associated with Parkinson's disease when production is
20    deficient).
21          The structures comprising the basal ganglia include but are not limited to: the putamen
22    and the globus  pallidus (together termed the lentiform nuclei), the pontine tegmentum, and the
23    caudate nuclei. Dystonia or involuntary muscle contraction can result from lesions in the
24    putamina; if there are concomitant lesions in the globus pallidus, Parkinsonism can result (Bhatia
25    and Marsden, 1994).  Bhatia and Marsden (1994) have discussed the various behavioral and
26    motor consequences of focal lesions of the basal ganglia from 240 case-study reports. Lesions in
27    the subcortical  white matter adjacent to the basal ganglia often occur as well (Airas et al., 2008;
28    Rubinstein et al., 1995; Bhatia and Marsden,  1994). In the case reports of Patankar et al. (1999),
29    it was noted that the severity and extent of necrosis in the lenticular nuclei do not necessarily
30    correlate with clinical outcome.

                                                4-1        DRAFT-DO NOT CITE OR QUOTE

-------
 1          In one of the earliest reviews of methanol overexposure, Bennett et al. (1953) described a
 2    mass accidental poisoning when 323 persons, ranging in age from 10 to 78 years, in Atlanta,
 3    Georgia, consumed "whisky" adulterated with as much as 35-40% methanol. In all, 41 people
 4    died. Of the 323 individuals, 115 were determined to be acidotic with symptoms (visual
 5    impairment, headache [affecting ~62%], dizziness [affecting ~30%], nausea, abdominal pain and
 6    others) beginning around 24 hours postexposure. Visual impairment was mostly characterized
 7    by blurred or indistinct vision; some who were not acidotic experienced transient visual
 8    disturbances.  The cardiovascular parameters were unremarkable.  The importance of acidosis to
 9    outcome is shown in Table 4-1. Among the key pathological features were cerebral edema, lung
10    congestion, gastritis, pancreatic necrosis, fatty liver, epicardial hemorrhages, and congestion of
11    abdominal viscera.
12          In another early investigation of methanol poisoning (involving 320 individuals), Benton
13    and Calhoun (1952) reported on methanol's visual disturbances.

          Table 4-1. Mortality rate for subjects exposed to methanol-tainted whiskey in relation
          to their level of acidosisaa
Subjects
All patients
Acidotic (CO2 <20 mEq)
Acidotic (CO2 <10 mEq)
Number
323
115
30
Percent deaths
6.2
19
50
      aThese data do not include those who died outside the hospital or who were moribund on arrival.
      Source: Bennett et al. (1953).

14          Riegel and Wolf (1966), in a case report involving a 60-year-old woman who ingested
15    methanol, noted that nausea and dizziness occurred within 30 minutes of ingestion.  She
16    subsequently passed out and remained unconscious for 3 days.  Upon awakening she had
17    paralysis of the vocal cords and was clinically blind in one eye after 4 months. Some aspects of
18    Parkinson-like symptoms were evident. There was a pronounced hypokinesia with a mask-like
19    face resembling a severe state of Parkinson's disease. The patient had difficulty walking and
20    could only make right turns with difficulty.  There was no memory loss.
21          Treatment of a 13-year-old girl who ingested an unspecified amount of a windshield-
22    washer solution containing 60% methanol was described by Guggenheim et al. (1971). She
23    displayed profound acidosis; her vital signs, once she was treated for acidosis, were normal by
24    36 hours after hospital admission.  During the ensuing 6 months after discharge from the
25    hospital, visual acuity (20/400, both eyes) worsened, and  she experienced muscle tremors, arm
26    pain, and difficulty in walking. A regimen of levadopa treatment greatly improved her ability to
27    function normally.

                                                4-2        DRAFT-DO NOT CITE OR QUOTE

-------
 1          Ley and Gali (1983) also noted symptoms that are Parkinson like following methanol
 2    intoxication.  In this case report respiratory support was needed; the woman was in a coma.
 3    Once stabilized, she exhibited symptoms similar to those noted in other case study reports, such
 4    as blurred vision, movement difficulty, and tremors. Computerized Axial Tomography scan
 5    findings highlighted the central nervous system (CNS) as an important site for methanol
 6    poisoning.
 7          Rubinstein et al. (1995) presented evidence that a methanol blood level of 36 mg/dL
 8    (360 mg/L) is associated with a suite of CNS and ocular deficits that led to a 36-year-old man
 9    (who subsequently died) becoming comatose.  CT scans at 1-2 days following ingestion were
10    normal. However, MRI scans at day 4 revealed lesions in the putamen and peripheral white
11    matter of the cerebral and cerebellar hemispheres.  Bilateral cerebellar cortical lesions had been
12    reported in an earlier case of methanol poisoning by Chen et al. (1991).
13          Finkelstein and Vardi (2002) reported that long-term inhalation exposure of a woman
14    scientist to methanol without acute intoxication resulted in a suite of delayed neurotoxic
15    symptoms (e.g., hand tremor, dystonia, bradykinesia, and other decrements in body movement).
16    Despite treatment with levadopa, an increase in the frequency and severity of effects occurred.
17    Exposure to bromine fumes was concomitant with exposure to methanol.
18          Hantson et al. (1997a) found, in four cases, that MRI and brain CT scans were important
19    tools in revealing specific brain lesions (e.g., in the putamina and white matter).  The first subject
20    was a 57-year-old woman who complained of blurred vision, diplopia, and weakness 24 hours
21    after ingesting 250 mL of a methanolic antifreeze solution. Upon hospital admission she was
22    comatose and in severe metabolic acidosis. An MRI scan at 9  days indicated abnormal
23    hyperintense foci in the putamina (decreased in size by day 23) and subtle lesions (no change by
24    day 23) in the white matter. Upon her discharge, bilateral deficits in visual acuity and color
25    discrimination persisted.
26          Similar deficits (metabolic acidosis, visual acuity, and color discrimination) were seen in
27    a man who ingested 300 mL of 75% methanol  solution. His blood methanol level was
28    163 mg/dL (1,630 mg/L).  An MRI administered 24 hours after hospital admission revealed
29    abnormal hyperintense foci in the putamina, with less intense lesions in the white matter.  Like
30    the first subject, a subsequent MRI indicated the foci decreased in size over time, but visual
31    impairments persisted.
32          The third individual, a male, ingested an unspecified amount of a methanolic solution.
33    His blood methanol level was 1,290 mg/dL (12,900 mg/L), and he was in a coma upon hospital
34    admission. An MRI revealed lesions in the putamina and occipital subcortical white matter.  A
35    follow-up CT scan was performed after 1 year and showed regression of the putaminal lesions


                                                4-3        DRAFT-DO NOT  CITE OR QUOTE

-------
 1   but no change in the occipital lesions. Upon his discharge, severe visual impairment remained
 2   but no extrapyramidal signs were observed.
 3          The last case was a man who became comatose 12 hours after ingesting 100 mL
 4   methanol.  His blood methanol level at that time was 60 mg/dL (600 mg/L). An MRI revealed
 5   lesions in the putamina;  at 3 weeks these lesions were observed to have decreased in size. Upon
 6   his discharge, the neurological signs had improved but optic neuropathy (in visual evoked
 7   potential) was observed.
 8          In a separate publication, Hantson et al. (1997b) reported a case of a 26-year-old woman
 9   who had ingested 250-500 mL methanol during the 38th week of pregnancy. Her initial blood
10   methanol level was 230 mg/dL (2,300 mg/L) (formate was 33.6 mg/dL or 336 mg/L), yet only a
11   mild metabolic acidosis was indicated. No distress to the fetus was observed upon gynecologic
12   examination.  Six days after therapy was initiated (methanol was not present in blood), she gave
13   birth. No further complications with either the mother or newborn were noted.
14          There have been several case reports involving  infant or toddler exposures to methanol
15   (De Brabander et al., 2005; Wu et al., 1995; Brent et al.,  1991; Kahn and Blum, 1979). The
16   report by Wu et al. (1995) involved a 5-week-old infant with moderate metabolic acidosis and a
17   serum methanol level of 1,148 mg/dL (11,480 mg/L), a level that is ordinarily fatal. However,
18   this infant exhibited no toxic signs and survived without any apparent permanent problems.  De
19   Brabander  et al. (2005) reported the case of a 3-year-old boy who ingested  an unknown amount
20   of pure methanol; at 3 hours after ingestion, the blood methanol level was almost 30 mg/dL (300
21   rng/L). Ethanol infusion as a therapeutic measure was  not well tolerated; at 8 hours after
22   ingestion, fomepizole was administered, and blood methanol levels stabilized below 20 mg/dL
23   (200 mg/L), a level above which is considered to be toxic by the American Academy of Clinical
24   Toxicology (Barceloux et al., 2002). Neither metabolic acidosis nor visual impairment was
25   observed in this individual.  Hantson et al. (1997a), in their review, touted the efficacy of
26   fomepizole over ethanol in the treatment of methanol poisoning
27          Bilateral putaminal lesions, suggestive of nonhemorrhagic necrosis  in the brain of a man
28   who accidentally ingested methanol, were reported by Arora et al. (2005).  Approximately
29   10 hours after MRI examination, he developed blurred vision and motor dysfunction. After
30   5 months, visual deficits persisted along with extrapyramidal symptoms. Persistent visual
31   dysfunction was also reported in another methanol poisoning case (Arora et al., 2007); the vision
32   problems developed -46 hours subsequent to the incident.
33          Vara-Castrodeza et al. (2007) applied diffusion-weighted MRI on a methanol-induced
34   comatose woman.  Diffusion-weighted MRI provides an image contrast distinct from standard
35   imaging  in that contrast is dependent on the molecular  motion of water (Schaefer et al., 2000).
36   The neuroradiological findings were suggestive of bilateral putaminal hemorrhagic necrosis,

                                                4-4        DRAFT-DO NOT CITE OR QUOTE

-------
 1    cerebral and intraventricular hemorrhage, diffuse cerebral edema, and cerebellar necrosis.
 2    Diffusion-weighted MRI allows for differentiation of restricted diffusion which is indicative of
 3    nonviable tissue.  In this case, treatment for acidosis (blood methanol levels had risen to
 4    1,000 mg/L) was unsuccessful and the patient died.
 5          Emergency treatment was unable to save the life of a 38-year-old man who presented
 6    with abdominal pain and convulsions after methanol intoxication (Henderson and Brubacher,
 7    2002).  A review of a head CT scan performed before the individual went into respiratory arrest
 8    revealed bilateral  globus pallidus ischemia.
 9          Discrete lesions of the putamen, cerebral white matter, and corpus callosum were
10    observed upon MRI (8 days post ingestion) in a man exposed to methanol (blood level 370
11    mg/L) complaining of vision loss (Keles et al., 2007).  Standard treatments corrected the acidosis
12    (pH 6.8), and at 1-month follow-up, his cognitive function improved but blindness and bilateral
13    optic atrophy were described as permanent. The follow-up MRI showed persistent putaminal
14    lesions with cortical involvement.
15          Fontenot and Pelak (2002) described a case of a woman who presented with persistent
16    blurred vision and a worsening mental status 36 hours after ingestion of an  unspecified amount
17    of methanol.  The initial CT scan revealed mild cerebral edema.  The blood methanol level at this
18    time was 86 mg/dL (860 mg/L). A repeat CT scan 48 hours after presentation showed
19    hypodensities in the putamen and peripheral white matter. One month after discharge, cognitive
20    function improved, and the patient experienced only a mild lower-extremity tremor.
21          Putaminal necrosis and edema of the deep white matter (the corpus  callosum was not
22    affected) was found upon MRI examination of a 50-year-old woman who apparently ingested an
23    unknown  amount of what was believed to be pure laboratory methanol (Kuteifan et al.,  1998).
24    Her blood methanol level was 39.7 mM (127 mg/dL; 1,272 mg/L) upon hospital admission and
25    dropped to 102 mg/dL (1,020 mg/L) at 10 hours and to 71 mg/dL (710 mg/L) at 34 hours. The
26    woman, a chronic alcoholic, was in a vegetative state when found and did not improved over the
27    course of a year.
28          MRI and CT scans performed on a 51-year-old man with generalized seizures who had a
29    blood methanol level of 95 mM (304 mg/dL; 3,044 mg/L) revealed bilateral hemorrhagic
30    necrosis of the putamen and caudate nuclei (Gaul et al., 1995). In addition, there was extensive
31    subcortical necrosis and bilateral necrosis of the pontine tegmentum and optic nerve. The patient
32    died several hours after the scans were performed.
33          The relation of methanol overexposure to brain hemorrhage was a focus of the report by
34    Phang et al. (1988), which followed the treatment of 7 individuals, 5 of whom died within
35    72 hours after hospital admission. In two of the deceased individuals, CT scans and autopsy
36    revealed putaminal hemorrhagic necrosis. The investigators postulated that the association of

                                               4-5        DRAFT-DO NOT CITE OR QUOTE

-------
 1    methanol with hemorrhagic necrosis may be complicated by the use of heparin during
 2    hemodialysis treatment for acidosis
 3          Treatment of two men who had drunk a solution containing 58% methanol and presented
 4    with impaired vision, coma, and seizures was discussed in a case report by Bessell-Browne and
 5    Bynevelt (2007).  A CT scan on one individual revealed bilateral putaminal and cerebral lesions.
 6    Blood methanol levels were 21 mg/L. This individual, despite standard treatments, never
 7    regained consciousness. The second individual, upon MRI, showed scattered hemorrhage at the
 8    grey-white interface of the cerebral hemispheres.
 9          There have been two case reports (Adanir et al., 2005; Downie et al., 1992) that involved
10    percutaneous and inhalation exposure.  Use of a methanol-containing emollient by a woman with
11    chronic pain led to vision loss, hyperventilation and finally, coma (Adanir et al., 2005).
12    Subsequent to standard treatment followed by hospital discharge, some visual impairment and
13    CNS decrements remained. The methanol blood threshold for ocular damage and acidosis
14    appeared to be -20 mg/L. Dutkiewicz et al. (1980) have determined the skin absorption rate to
15    be 0.192 mg/cm2/minute.  In the case report of Aufderheide et al. (1993), two firefighters were
16    transiently exposed to methanol by inhalation and the percutaneous route.  Both only complained
17    of a mild headache and had blood methanol levels of 23 and 16 mg/dL (230 and 160 mg/L),
18    respectively.
19          Bebarta et al. (2006) conducted a prospective observational study of seven men who had
20    purposefully inhaled a methanol-containing product. Four had a blood methanol level upon
21    hospital presentation of >24 mg/dL (240 mg/L); the mean formic acid level was 71 |ig/dL.  One
22    individual had a blood methanol level of 86 mg/dL (860 mg/L) and a blood formic acid level of
23    250 |ig/mL upon hospital admission. This latter individual was treated with fomepizole. No
24    patient had an abnormal ophthalmologic examination.  All seven stabilized quickly and acidosis
25    was normalized in 4 hours.
26          Numerous other case reports documenting putaminal necrosis/hemorrhage and/or
27    blindness have been reported (Blanco et al., 2006; Feany et al., 2001; Hsu et al., 1997; Pelletier
28    etal., 1992; Chen etal., 1991).
29          Hovda et al. (2005) presented a combined prospective and retrospective case series study
30    of 51  individuals in Norway (39 males and 12 females, many  of whom were alcoholics) who
31    were hospitalized after  consuming tainted spirits containing 20% methanol and 80% ethanol.  In
32    general, serum methanol concentrations were highest among those most severely affected.  The
33    poor outcome was closely correlated with the degree of metabolic acidosis. It was noted by the
34    investigators that the concomitant consumption of ethanol prevented more serious sequelae in
35    2/5 individuals who presented with detectable ethanol levels and were not acidotic despite 2
36    having the highest blood methanol levels. However, others with detectable levels of ethanol

                                               4-6        DRAFT-DO NOT CITE OR QUOTE

-------
 1    along with severe metabolic acidosis (two of whom died) presumably had subtherapeutic levels
 2    of ethanol in their system.
 3          In a later report, Hovda et al. (2007) focused on formate kinetics in a 63-year-old male
 4    who died 6 days after being admitted to the hospital with headache, vomiting, reduced vision,
 5    and dizziness. The investigators speculated that the prolonged metabolic acidosis observed (T1/2
 6    for formic acid was 77 hours before dialysis, compared to a typical normal range of 2.5-
 7    12 hours) may have been related to retarded formate elimination.
 8          Hovda and colleagues (Hunderi et al., 2006) found a strong correlation between blood
 9    methanol concentration and the osmolal gap (R2 = 0.92) among 17 patients undergoing dialysis
10    after consuming methanol-contaminated spirits. They concluded that the osmolal gap could be
11    taken as a priori indication of methanol poisoning and be used to guide initiation and duration of
12    dialysis. As they indicated, many hours of dialysis could be safely dispensed with.  The osmolal
13    gap pertains to the effect that methanol (and other alcohols) has on the depression of the freezing
14    point of blood in the presence of normal solutes. Braden et al. (1993) demonstrated in case
15    studies that the disappearance of the osmolal gap correlates with the correction of acidosis; they
16    cautioned that methanol and ethanol should not be assumed to be the main factors in causing
17    osmolal gap as glycerol and acetone and its metabolites can as well.  A more detailed discussion
18    of the anion and osmolal gap has been provided by Henderson and Brubacher (2002).
19          Hassanian-Moghaddam et al. (2007) compiled data on the prognostic factor relating to
20    outcome in methanol-poisoning cases in Iran.  They examined 25 patients, 12 of whom died; 3
21    of the survivors were rendered blind.  There was a significant difference in mean pH of the first
22    arterial blood gas measurements of those who subsequently died compared with survivors. It
23    was concluded that poor prognosis was associated with pH <7, coma upon admission, and
24    >24-hours delay from intake to admission.
25          The use of blood methanol levels as predictors of outcome is generally not recommended
26    (Barceloux et al., 2002).  These investigators cited differences in sampling time, ingestion of
27    ethanol, and levels of toxic (e.g., formic acid) metabolites among the complicating factors. As
28    an illustration, the case report by Prabhakaran et al. (1993) cites two women who ingested a
29    methanol solution (photocopying diluent) at about the same time, were admitted to the hospital
30    about the same time (25-26 hours after ingestion) and had identical plasma methanol
31    concentrations (83 mg/dL; 830 mg/L) upon admission, but different outcomes. Patient #1 was in
32    metabolic acidosis and had an unstable conscious  state even after treatment. Upon discharge at
33    day 6, there were no apparent sequelae. Patient #2 had severe metabolic acidosis, fixed and
34    dilated pupils, and no brain stem reflexes.  This patient died at day 3 even though therapeutic
35    measures had been administered.
                                                4-7        DRAFT-DO NOT CITE OR QUOTE

-------
 1          In a discussion of 3 fatal methanol-overexposure cases, Andresen et al. (2008) found
 2    antemortem blood methanol levels of 540 and 740 mg/dL (5,400 and 7,400 mg/L) in 2
 3    individuals. At autopsy brain stem blood levels were 738 and 1,008 mg/dL (7,380 and
 4    10,080 mg/L), respectively. These brain levels were much higher than blood levels postmortem.
 5    Autopsy revealed brain and pulmonary edema in all three individuals; in the two who had the
 6    longer survival times, there was hemorrhagic necrosis of the putamen and hemorrhages of the
 7    tissue surrounding the optic nerve. In their study of 26 chronic users of methylated spirits,
 8    Meyer et al. (2000) found that the best predictor of death or a poor outcome in chronic abusers
 9    was a pH <7.0; there was no correlation between blood methanol levels and outcome. Mahieu
10    et al. (1989) considered a latency period before treatment exceeding 10 hours and a blood
11    formate level >50 mg/dL (500 mg/L) as predictive of possible permanent sequelae. Liu et al.
12    (1998) in their examination of medical records of 50 patients treated for methanol poisoning over
13    a 10-year period found that 1) deceased patients had a higher mean blood methanol level than
14    survivors; and 2) initial arterial pH levels <7.0 (i.e., severe metabolic acidosis). Coma or seizure
15    was also associated with higher mortality upon hospital admission.
16          Numerous cases of methanol poisoning have been documented in a variety of countries.
17    In Tunisia, 16 cases of methanol poisoning were discussed by Brahmi et al. (2007).  Irreversible
18    blindness occurred in two individuals, with others reporting CNS symptoms, GI effects, visual
19    disturbances, and acidosis. Putaminal necrosis was also described in case reports from Iran
20    (Sefidbakht et al., 2007). Of 634 forensic autopsies carried out in Turkey  during 1992-2003, 18
21    appeared to be related to methanol poisoning (Azmak, 2006). Brain edema and focal necrosis of
22    the optic nerve were  among various sequelae noted.  Dethlefs and colleagues (Naraqi et al.,
23    1979; Dethlefs and Naraqi, 1978) described permanent ocular damage in 8/24 males who
24    ingested methanol in Papua New Guinea.
25          In summary, most cases of accidental/intentional methanol poisoning reveal a common
26    set of symptoms, many of which are likely to be presented upon hospital admission. These
27    include:
28              •   blurred vision and bilateral or unilateral blindness
29              •   convulsions, tremors, and coma
30              •   nausea, headache, and dizziness
31              •   abdominal pain
32              •   diminished motor skills
33              •   acidosis
34              •   dyspnea
35              •   behavioral and/or emotional deficits
36              •   speech impediments
                                               4-8        DRAFT-DO NOT CITE OR QUOTE

-------
 1          Acute symptoms generally are nausea, dizziness, and headache. In the case reports cited
 2    above, the onset of symptom sets as well as their severity varies depending upon how much
 3    methanol was ingested, whether or not and when appropriate treatment was administered, and
 4    individual variability. A longer time between exposure and treatment, with few exceptions,
 5    results in more severe outcomes (e.g., convulsions, coma, blindness, and death). The diminution
 6    of some acute and/or delayed symptoms may reflect concomitant ingestion of ethanol or how
 7    quickly therapeutic measures (one of which includes ethanol infusion) were administered in the
 8    hospital setting.
 9          Those individuals who are in a metabolic acidotic state (e.g., pH <7.0) are typically the
10    individuals who manifest the more severe symptoms. Many case reports stress that, unlike blood
11    pH levels <7.0, blood levels of methanol are not particularly good predictors of health outcome.
12    According to a publication of the American Academy of Clinical Toxicology (Barceloux et al.,
13    2002), "the degree of acidosis at presentation most consistently correlates with  severity and
14    outcome."
15          As the case reports demonstrate, those individuals who present with more severe
16    symptoms (e.g.,  coma, seizures, severe acidosis) generally exhibit higher mortality (even after
17    treatment) than those without such symptoms.  In survivors of poisoning, persistence or
18    permanence of vision decrements and particularly blindness often have been observed
19          Correlation of symptomatology with blood levels of methanol has been  shown to vary
20    appreciably between individuals. Blood methanol levels in the case reports involving ingestion
21    ranged from values of 30 to over 1,000 mg/dL (300 to over 10,000 mg/L).  The lowest value
22    (20 mg/dL; 200 mg/L) reported (Adanir et al., 2005) involved a case of percutaneous absorption
23    (with perhaps associated inhalation exposure) that led to vision and CNS deficits after hospital
24    discharge. In one case report (Rubinstein et al., 1995) involving ingestion, coma and subsequent
25    death were associated with an initial blood methanol level of 36 mg/dL (360 mg/L).
26          Upon MRI and CT scans, the more seriously affected individuals typically have focal
27    necrosis in both brain white matter and more commonly, in the putamen.  Bilateral hemorrhagic
28    and nonhemorrhagic necrosis of the putamen is considered by many radiologists as the most
29    well-known sequelae of methanol overexposure.

      4.1.2. Occupational Studies
30          Occupational health studies have been carried out to investigate the potential effects of
31    chronic exposure to lower levels of methanol than those seen in acute poisoning cases such as
32    those described above. For example, Frederick et al. (1984) conducted a health hazard
33    evaluation on behalf of the National Institute for Occupational Safety and Health (NIOSH) to
                                               4-9         DRAFT-DO NOT CITE OR QUOTE

-------
 1    determine if vapor from duplicating fluid (which contains 99% methanol) used in mimeograph
 2    duplicating machines caused adverse health effects in exposed persons. A group of 84 teacher's
 3    aides were selected for study, 66 of whom responded with a completed medical questionnaire. A
 4    group of 297 teachers (who were not exposed to methanol vapors to the same extent as the
 5    teacher's aides) completed questionnaires as a control group.  A 15-minute breathing zone
 6    sample was taken from 21 duplicators, 15 of which were greater than the NIOSH-recommended
 7    short term ceiling concentration of 800 ppm (1048 mg/m3). The highest breathing zone
 8    concentrations were in the vicinity of duplicators for which no exhaust ventilation had been
 9    provided (3,080 ppm [4,036 mg/m3] was the highest value recorded). Upon comparison of the
10    self-described symptoms of the 66 teacher's aides with those of 66 age-matched teachers chosen
11    from the 297 who responded, the number of symptoms potentially related to methanol were
12    significantly higher in the teacher's aides. These included blurred vision (22.7 versus 1.5%),
13    headache (34.8 versus 18.1%), dizziness (30.3 versus  1.5%), and nausea (18 versus 6%). By
14    contrast, symptoms that are not usually associated with methanol exposure (painful urination,
15    diarrhea, poor appetite, and jaundice) were similar in incidence among the groups.
16          To further investigate these disparities, NIOSH physicians (not involved in the study)
17    defined a hypothetical case of methanol toxicity by any of the following four symptom
18    aggregations: 1) visual  changes; 2) one acute symptom (headache, dizziness, numbness,
19    giddiness, nausea or vomiting) combined with one chronic symptom (unusual fatigue, muscle
20    weakness, trouble sleeping, irritability, or poor memory); 3) two acute symptoms; or 4) three
21    chronic symptoms. By these criteria, 45% of the teacher's aides were classified as  being
22    adversely affected by methanol exposure compared to 24% of teachers (p < 0.025).  Those
23    teacher's aides and teachers who spent a greater amount of time using the duplicators were
24    affected at a higher rate than  those who used the machines for a lower percentage of their work
25    day.
26          Tanner (1992) reviewed the occupational and environmental causes of Parkinsonism,
27    spotlighting the potential etiological significance of manganese, carbon monoxide,  repeated head
28    trauma (such as suffered by boxers), and exposure to solvents. Among the latter, Tanner (1992)
29    discussed the effects of methanol and n-hexane on the nervous system.  Acute methanol
30    intoxication resulted in inebriation, followed within hours by GI pain, delirium, and coma.
31    Tanner (1992) pinpointed the formation of formic acid, with consequent inhibition  of
32    cytochrome oxidase, impaired mitochondrial function, and decreased ATP formation as relevant
33    biochemical and physiological changes for methanol exposure.  Nervous system injury usually
34    includes blindness, Parkinson-like symptoms, dystonia, and cognitive impairment,  with injury to
35    putaminal neurons most likely underlying the neurological responses.


                                               4-10        DRAFT-DO NOT CITE OR QUOTE

-------
 1          Kawai et al. (1991) carried out a biomarker study in which 33 occupationally exposed
 2   workers in a factory making methanol fuel were exposed to concentrations of methanol of up to
 3   3,577 ppm (4,687 mg/m3), as measured by personal samplers of breathing zone air.  Breathing
 4   zone exposure samples were correlated with the concentrations of methanol in urine at the end of
 5   the shift in 38 exposed individuals and 30 controls (r = 0.82). Eleven of 22 individuals who
 6   experienced high exposure to methanol (geometric mean of 459 ppm [601 mg/m3]) complained
 7   of dimmed vision during work while 32% of this group of workers experienced nasal irritation.
 8   These incidences were statistically significant compared to those of persons who worked in low-
 9   exposure conditions (geometric mean of 31 ppm [41 mg/m3]). One 38-year-old female worker
10   who had worked at the factory for only 4 months reported that her visual acuity had undergone a
11   gradual impairment.  She also displayed a delayed light reflex.
12          Lorente et al. (2000) carried out a case control study of 100 mothers whose babies had
13   been born with cleft palates.  Since all of the mothers had worked during the first trimester,
14   Lorente et al. (2000) examined the occupational information for each subject in comparison to
15   751 mothers whose babies were healthy. Industrial hygienists analyzed the work histories of all
16   subjects to determine what, if any, chemicals the affected mothers may have been exposed to
17   during pregnancy.  Multivariate analysis was used to calculate odds ratios, with adjustments
18   made for center of recruitment, maternal age, urbanization, socioeconomic status, and country of
19   origin.  Occupations with positive outcomes for cleft palate in the progeny were hairdressing
20   (OR = 5.1, with a 95% confidence interval [CI] of 1.0-26) and housekeeping (OR = 2.8, with a
21   95% CI of 1.1-7.2).  Odds ratios for cleft palate only and cleft lip with or without cleft palate
22   were calculated for 96 chemicals. There seemed to be no consistent pattern of association for
23   any chemical or group of chemicals with these impairments, and possible exposure to methanol
24   was negative for both outcomes.

     4.1.3. Controlled Studies
25          Two controlled studies have evaluated humans for neurobehavioral function following
26   exposure to -200 ppm (262 mg/m3) methanol  vapors in a controlled setting.  The occupational
27   TLV established by the American Conference  of Governmental Industrial Hygienists (ACGIH,
28   2000) is 200 ppm (262 mg/m3). In a pilot study by Cook et al. (1991), 12 healthy young men
29   (22-32 years of age) served as their own controls and were tested for neurobehavioral function
30   following a random acute exposure to air or 191 ppm (250 mg/m3) methanol vapors for
31   75 minutes. The majority of results in a battery of neurobehavioral endpoints were negative.
32   However, statistical significance was obtained for results in the P-200 and N1-P2 component of
33   event-related potentials (brain wave patterns following light flashes and sounds), the Sternberg
34   memory task, and subjective evaluations of concentration and fatigue. As noted by the Cook et
                                               4-11       DRAFT-DO NOT CITE OR QUOTE

-------
 1    al. (1991), effects were mild and within normal ranges. Cook et al. (1991) acknowledged
 2    limitations in their study design, such as small sample size, exposure to only one concentration
 3    for a single duration time, and difficulties in masking the methanol odor from experimental
 4    personnel and study subjects.
 5          In a randomized double-blind study, neurobehavioral testing was conducted on 15 men
 6    and 11 women (healthy, aged 26-51 years) following exposure to 200 ppm (262 mg/m3)
 7    methanol or water vapors for 4 hours (Chuwers et al.,  1995); subjects served as their own
 8    controls in this study.  Exposure resulted in elevated blood and urine methanol levels (up to peak
 9    levels of 6.5 mg/L and 0.9 mg/L, respectively) but not formate concentrations. The majority of
10    study results were negative.  No significant findings were noted for visual, neurophysiological,
11    or neurobehavioral tests except  for slight effects (p < 0.05) on P-300 amplitude (brain waves
12    following exposure to sensory stimuli) and Symbol Digit testing (ability to process information
13    and psychomotor skills). Neurobehavioral performance was minimally affected by methanol
14    exposure at this level. Limitations noted by Chuwers et al. (1995) are that studies of alcohol's
15    affect on P-300 amplitude suggest that this endpoint may be biased by unknown factors and
16    some experimenters and subjects correctly guessed if methanol was used.
17          Although the slight changes in P-200 and P-300 amplitude noted in both the Chuwers
18    et al. (1995) and Cook et al. (1991) studies may be an  indication of moderate alterations in
19    cognitive function, the results of these studies are generally consistent and suggest that the
20    exposure concentrations employed were below the threshold for substantial neurological effects.
21    This is consistent with the data from acute poisoning events which have pointed to a serum
22    methanol threshold of 200 mg/L for the instigation of acidosis, visual impairment, and CNS
23    deficits.
24          Mann et al. (2002) studied the effects of methanol exposure on human respiratory
25    epithelium as manifested by local irritation, ciliary function, and immunological factors. Twelve
26    healthy men (average age 26.8 years) were exposed to 20 and 200 ppm (26.2 and 262 mg/m3,
27    respectively) methanol for 4 hours at each concentration; exposures were separated by 1-week
28    intervals. The 20 ppm (26.2 mg/m3) concentration was considered to be the  control exposure
29    since previous studies had demonstrated that subjects can detect methanol concentrations of
30    20 ppm (26.2 mg/m3) and greater. Following each single exposure, subclinical inflammation was
31    assessed by measuring concentrations of interleukins (IL-8, IL-lp, and IL-6) and prostaglandin
32    E2 in nasal secretions.  Mucociliary clearance was evaluated by conducting a saccharin transport
33    time test and measuring ciliary beat frequency.  Interleukin and prostaglandin data were
34    evaluated by a 1-tailed Wilcoxon test, and ciliary function data were assessed by a 2-tailed
35    Wilcoxon test. Exposure to 200 (262 mg/m3) versus 20 ppm (26.2 mg/m3) methanol resulted in a
36    statistically-significant increase in IL-lp (median of 21.4 versus 8.3 pg/mL)  and IL-8 (median of

                                               4-12        DRAFT-DO NOT CITE OR QUOTE

-------
 1    424 versus 356 pg/mL).  There were no significant effects on IL-6 and prostaglandin E2
 2    concentration, ciliary function, or on the self-reported incidence of subjective symptoms of
 3    irritation. The authors concluded that exposure to 200 ppm (262 mg/m3) methanol resulted in a
 4    subclinical inflammatory response.
 5          In summary, adult human subjects acutely exposed to 200 ppm (262 mg/m3) methanol
 6    have experienced slight neurological (Chuwers et al., 1995) and immunological effects
 7    (increased subclinical biomarkers for inflammation) with no self-reported symptoms of irritation
 8    (Mann et al., 2002). These exposure levels were associated with peak methanol blood levels of
 9    6.5 mg/L (Chuwers et al., 1995), which is approximately threefold higher than background
10    methanol blood levels reported for adult human subjects on methanol-restrictive diets
11    (Table 3-1). Nasal irritation effects have been reported by adult workers exposed to 459 ppm
12    (601 mg/m3) methanol (Kawai et al., 1991).  Frank effects such as blurred vision, bilateral or
13    unilateral blindness, coma, convulsions/tremors, nausea, headache, abdominal pain,  diminished
14    motor skills, acidosis, and dyspnea begin to occur as blood levels approach 200 mg methanol/L,
15    while 800 mg/L appears to be the threshold for lethality. Data for subchronic, chronic or in utero
16    human exposures are very limited and inconclusive.

      4.2. ACUTE, SUBCHRONIC AND CHRONIC STUDIES AND CANCER BIO ASSAYS IN
      ANIMALS—ORAL AND INHALATION
17          A number of studies in animals have  investigated the acute, subchronic, and  chronic
18    toxicity of methanol. Most are via the inhalation route.  Presented below are summaries of these
19    investigations.

      4.2.1.  Oral Studies
      4.2.1.1. Acute  Toxicity
20          Although there are few studies that have examined the short-term toxic effects  of
21    methanol via the oral route, a number of median lethal dose (LD50) values have been published
22    for the compound. As listed in Lewis (1992), these include 5,628 mg/kg in rats, 7,300 mg/kg in
23    mice, and 7,000 mg/kg in monkeys.
      4.2.1.2. Subchronic Toxicity
24          An oral repeat dose study was conducted by the EPA (1986c) in rats. Sprague-Dawley
25    rats (30/sex/dose) were gavaged with 0, 100, 500, or 2,500 mg/kg-day of methanol.  Six weeks
26    after dosing, 10 rats/sex/dose group were subjected to interim sacrifice, while the remaining rats
27    continued on the dosing regimen until the final sacrifice (90 days). This  study generated data on
28    weekly body weights and food consumption, clinical signs of toxi city, ophthalmologic
29    evaluations, mortality, blood and urine chemistry (from  a comprehensive set of hematology,
                                               4-13       DRAFT-DO  NOT CITE OR QUOTE

-------
 1    serum chemistry, and urinalysis tests), and gross and microscopic evaluations for all test animals.
 2    Complete histopathologic examinations of over 30 organ tissues were done on the control and
 3    high-dose rats.  Histopathologic examinations of livers, hearts, and kidneys and all gross lesions
 4    seen at necropsy were done on low-dose and mid-dose rats. There were no differences between
 5    dosed animals and controls in body weight gain, food  consumption, or upon gross or microscopic
 6    evaluations. Elevated levels (p < 0.05 in males) of serum alanine transaminase (ALT)14 and
 7    serum alkaline phosphatase (SAP), and increased (but not statistically significant) liver weights
 8    in both male and female rats suggest possible treatment-related effects in rats bolus dosed with
 9    2,500 mg methanol/kg-day despite the absence of supportive histopathologic lesions in the liver.
10    Brain weights of high-dose group (2,500 mg/kg-day) males and females were significantly less
11    than those of the control group at terminal sacrifice. Based on these findings, 500 mg/kg-day of
12    methanol is considered an NOEL from this rat study.
      4.2.1.3. Chronic Toxicity
13           A report by Soffritti et al. (2002a) summarized a European Ramazzini Foundation (ERF)
14    chronic duration experimental study of methanol15 in which the compound was provided to
15    100 Sprague-Dawley rats/sex/group ad libitum in drinking water at concentrations of 0, 500,
16    5,000, and 20,000 ppm (v/v). The animals were 8 weeks old at the beginning of the study. In
17    general, ERF does not randomly assign animals to treatment groups, but assigns all animals from
18    a given litter to the same treatment group (Bucher, 2002). All rats were exposed for up to
19    104 weeks, then maintained until they died naturally.  Rats were housed in groups of 5 in
20    Makrolon cages (41  x 25 x 15  cm) in a room that was maintained at 23 ± 2°C and 50-60%
21    relative humidity. The in-life portion of the experiment ended at 153 weeks with the death of the
22    last animal. Mean daily drinking water, food consumption, and body weights were monitored
23    weekly for the first 13 weeks, every 2 weeks thereafter for 104 weeks, then every  8 weeks until
24    the end of the experiment. Clinical signs were monitored 3 times/day, and the occurrence of
25    gross changes was evaluated every 2 weeks. All rats were necropsied at death then underwent
26    histopathologic examination of organs and tissues.16
      14 Also known as serum glutamate pyruvate transaminase (SGPT)
      15 Soffritti et al. (2002a) report that methanol was obtained from J.T. Baker, Deventer, Holland, purity grade 99.8%.
        Histopathology was performed on the following organs and tissues: skin and subcutaneous tissue, brain, pituitary
      gland, Zymbal glands, parotid glands, submaxillary glands, Harderian glands, cranium (with oral and nasal cavities
      and external and internal ear ducts) (5 sections of head), tongue, thyroid and parathyroid, pharynx, larynx, thymus
      and mediastinal lymph nodes, trachea, lung and mainstem bronchi, heart, diaphragm, liver, spleen, pancreas,
      kidneys, adrenal glands, esophagus, stomach (fore and glandular), intestine (four levels), urinary bladder, prostate,
      gonads, interscapular fat pad, subcutaneous and mesenteric lymph nodes, and any other organs or tissues with
      pathologic lesions.
                                                 4-14         DRAFT-DO NOT CITE OR QUOTE

-------
 1          Soffritti et al. (2002a) reported no substantial dose-related differences in survival, but no
 2    data were provided. Using individual animal data available from the ERF website,17 Cruzan
 3    (2009) reports that male rats treated with methanol generally survived better than controls, with
 4    50% survival occurring at day 629, 686, 639 and 701 in the 0, 500, 5,000, and 20, 000 mg/L
 5    groups, respectively. There were no significant differences in survival between female control
 6    and treatment groups, with 50% survival occurring at day 717, 691, 678 and 708 in the 0,  500,
 7    5,000, and 20, 000 mg/L groups, respectively. Body weight and water and food consumption
 8    were monitored in the study, but the data were not documented in the published report.
 9    However, based on data available from the ERF website, average doses of 0, 53.2, 524, and
10    1,780 mg/kg-day in males and 0, 66.0, 624.1, and 2,177 mg/kg-day in females could be
11    calculated (see Appendix E) from drinking water concentrations of 0, 500, 5,000, and
12    20,000 ppm.
13          Soffritti et al. (2002a) reported that water consumption in high-dose females was reduced
14    compared to controls between 8 and 56 weeks and that the mean body weight in high-dose males
15    tended to be higher than that of control males. Overall, there was no pattern of compound-
16    related clinical signs of toxicity, and the available data did not provide any indication that the
17    control group was not concurrent with the treated group (Cruzan, 2009).  Soffritti et al. (2002a)
18    further reported that there were no compound-related signs of gross pathology or histopathologic
19    lesions indicative of noncancer toxicological  effects in response to methanol.
20          Soffritti et al. (2002a) reported a number of oncogenic responses to methanol (Table 4-2),
21    principally hemolymphoreticular neoplasms,  the majority of which were reported to be lympho-
22    immunoblastic lymphomas.  In ERF bioassays, including this methanol study,
23    hemolymphoreticular neoplasms are generally divided into specific histological types
24    (lymphoblastic lymphoma, lymphoblastic leukemia, lymphocytic lymphoma, lympho-
25    immunoblastic lymphoma, myeloid leukemia, histocytic sarcoma, and monocytic leukemia) for
26    identification purposes. According to Soffritti et al. (2007), the overall incidence of
27    hemolymphoreticular tumors (lymphomas/leukemias) in ERF studies is 13.3% (range, 4.0-
28    25.0%) in female historical controls (2,274 rats) and 20.6% (range, 8.0-30.9%) in male historical
29    controls (2,265 rats).  The high-dose responses, shown in Table 4-2, of 28% and 40% for females
30    and males, respectively, are above their corresponding historical ranges.18
31          The National Toxicology Program (NTP) does not routinely subdivide lymphomas into
32    specific histological types as was done by the ERF.  In 2004, a Pathology Working Group
33    (PWG) of National Institute of Environmental Health Sciences (NIEHS) performed a limited
34    review of about 75 slides provided by ERF as representive of lesions in Sprague-Dawley rats
      17 http://www.ramazzini.it/fondazione/foundation.asp.
      18 While historical control data can be informative, for reasonably well-conducted studies, it should not take
      precedence over concurrent controls or appropriate statistical dose-response trend tests.
                                               4-15        DRAFT-DO NOT CITE OR QUOTE

-------
 1    associated with aspartame exposure (EPS A, 2006; Hailey, 2004). The primary objective of this
 2    review was to "provide a second opinion for this set of lesions by a group of pathologists
 3    experienced in Toxicologic Pathology."19 Eleven of the slides reviewed by the PWG were
 4    related to lymphomas, and three of these had been classified by ERF as lympho-immunoblastic.
 5    The PWG concluded that "The diagnoses of lymphatic and histocytic neoplasms in the cases
 6    reviewed were generally confirmed" (Hailey, 2004).  In particular, the PWG accepted the more
 7    specific diagnoses of ERF when the lesions were considered to be consistent with a neoplasm of
 8    lymphocytic, histocytic, monocytic, and/or myeloid origin. The PWG noted, however, that while
 9    lymphoblastic lymphomas, lymphocytic lymphomas, lympho-immunoblastic lymphomas, and
10    lymphoblastic leukemias as malignant lymphomas can be combined, myeloid leukemias,
11    histocytic sarcomas, and monocytic leukemia should be treated as separate malignancies and not
12    combined with the other lymphomas since they are of different cellular origin (Hailey, 2004).
13    McConnell et al. (1986) and Cruzan (2009) have also noted that myeloid leukemia, histocytic
14    sarcoma, and monocytic leukemia are of a different cell line and are not typically combined with
15    other lymphomas for statistical significance or dose-response modeling. Consistent with these
16    judgments, EPA has not included the myeloid leukemia, histocytic sarcoma, and monocytic
17    leukemia in combination with lymphoblastic lymphoma, lymphoblastic leukemia, lymphocytic
18    lymphoma, and lympho-immunoblastic lymphoma in its consideration of tumorgenic responses
19    reported by ERF (see Section 5.4.1; Table 5-6). Thus, EPA's analysis of this tumorogenic
20    response differs from the lymphoreticular tumor response shown in Table 4-2 and reported by
21    Soffriti etal. (2002a). As described in Section 5.4.1.1, EPA's reassessment indicates a
22    significant increase in tumor response at the two highest doses for males and across all doses for
23    females (Fisher's exact,p < 0.05), as well as a significant dose-response trend (Cochran
24    Armitage trend test; p < 0.05).
25          Schoeb et al. (2009) have suggested that the interpretation of lesions in ERF studies,
26    including the Soffritti et al. (2002a) methanol study, may have been confounded by a respiratory
27    infection referred to as Mycoplasma pulmonis (M. pulmonis) disease and that lesions of this
28    disease were interpreted as lymphoma.  They noted that lympho-immunoblastic lymphoma is not
29    listed as a lymphoma type in rats in available reference sources and that the cellular morphology
30    of the lung lympho-immunoblastic lymphomas reported by ERF for aspartame  (Soffritti et al.,
31    2005) and MTBE (Belpoggi et al., 1999) studies are more consistent withM pulmonis disease.
32    As noted above, an NIEHS PWG (Hailey, 2004) has confirmed the ERF diagnosis of the several
33    lymphomas, including three lymphomas from the lung, thymus and medullary lymph node and
34    mesenteric lymph node that were characterized by ERF as "lympho-immunoblastic."  Hailey
      19 This review was not considered a "peer review" of the pathology data from this study. As noted by Hailey (2004),
      "a peer review would necessitate a review of the study data by a second party, and selection and examination of
      lesions based upon that data review."
                                              4-16        DRAFT-DO NOT CITE OR QUOTE

-------
 1    (2004) reports that the PWG "accepted their [ERF's] more specific diagnosis if the lesion was
 2    considered to be consistent with a neoplasm of lymphocytic, histocytic, monocytic, and/or
 3    myeloid origin."  The concerns of Schoeb et al. (2009) regarding the possibility of infection
 4    confounding the interpretation of lung lesions in the ERF study are not unfounded. Chronic
 5    inflammatory changes are apparently a common finding in ERF studies (Caldwell et al., 2008),
 6    probably caused by the ERF bioassay design that does not employ specific pathogen-free (SPF)
 7    rats (EFSA, 2006) and allows the rats to live out their "natural life span" in the absence of
 8    disease barriers (e.g., fully enclosed cages). However, an infection of the ERF colony with
 9    M. pulmonis has not been confirmed (Caldwell et al., 2008) and, without confirmation, cannot be
10    used to discount an existing dose-related trend (U.S. EPA, 2005).  Further, even if the rats of the
11    ERF methanol study were suffering from a respiratory infection that confounded the
12    interpretation of lung lesions, 60% of reported lymphoma incidences involved other organ
13    systems, and the dose-response for lymphomas in other organ systems is not remarkably
14    different than for all lymphomas (see analysis in Section 5.4.3.2).
15          Another cancer response, reported by Soffritti et al. (2002a), that is considered to be
16    potentially related to methanol exposure was an increase in rare hepatocellular carcinomas in
17    male rats. Although the increase was not statistically increased compared to concurrent controls,
18    EPA has analyzed historical data for this tumor type in this species and  determined that the
19    incidence in all dose groups was significantly  elevated relative to historical controls (Fisher's
20    exact/? < 0.05 for all doses and/? < 0.01 for the high-dose group).  The historical control group
21    (n = 407) used was the combined control groups from ERF studies for which individual animal
22    pathology data have been made available via the ERF website20 and include data for methanol,
23    formaldehyde, aspartame, MTBE, and TAME.
24          As noted in Table 4-2, increased incidences of carcinomas of the ear ducts and
25    osteosarcomas of the head were reported for both female and male rats, with a statistically
26    significant increase in only the high-dose male ear duct carcinomas. Ear duct carcinomas are a
27    rare finding in Charles River rats and NTP historical  databases of Sprague-Dawley rats (Cruzan,
28    2009). In their limited review of pathology slides from the ERF aspartame bioassay (Soffritti
29    et al., 2006, 2005), NTP pathologists interpreted a majority of such head pathologies, including
30    in the ear duct, as being hyperplastic in nature, not carcinogenic (EFSA, 2006; Hailey, 2004).
31    Soffritti  et al. (2002a) also noted an increased  incidence of testicular hyperplasia in high-dose
32    males  and uterine sarcomas in high-dose females compared to controls. However, these
33    increases were not statistically significant and were within historical control ranges for this
34    species and strain (NTP, 2007, 1999; Haseman et al., 1998).  The group-specific total number of
35    malignant tumors was also shown to increase with dose in both sexes of rats.
      20 http://www.ramazzini.it/fondazione/foundation. asp.
                                                4-17       DRAFT-DO NOT CITE OR QUOTE

-------
Table 4-2. Incidence of carcinogenic
methanol in drinking water for up to
                                             responses in Sprague-Dawley rats exposed to
                                             2 years
Tissues/affected sites
Ear duct (carcinomas)
Head (osteosarcomas)
Hemolymphoreticular tumors
Liver (hepatocarcinomas)
Testis (interstitial cell adenomas)
Total malignant tumors
Dose (mg/kg-day)
Males
0
9/100
6/100
28/100
0/100
12/100
50/100
53.2
13/100
6/100
35/100
2/100
9/100
55/100
524
17/100
13/100
36/100
2/100
13/100
64/100
1780
24/100b
11/100
40/100
3/100
17/100
70/100b
Females
0
9/100
1/100
13/100
0/100

43/100
66.0
8/100
4/100
24/100
0/100

48/100
624.1
16/100
3/100
24/100
1/100

48/100
2177
19/100
6/100
28/1003
0/100

63/100b
             <_;
      "p < 0.05 using the ^ test.
      V < 0.01 using the -£ test.
      Source: Soffritti et al. (2002a).
 1          Apaja (1980) performed dermal and drinking water chronic bioassays in which male and
 2    female Eppley Swiss Webster mice (25/sex/dose group; 8 weeks old at study initiation) were
 3    exposed 6 days per week until natural death to various concentrations of malonaldehyde and
 4    methanol. The stated purpose of the study was to determine the carcinogen! city of
 5    malonaldehyde, a product of oxidative lipid deterioration in rancid beef and other food products
 6    in advanced stages of degradation. However, due to its instability, malonaldehyde was  obtained
 7    from the more stable malonaldehyde bis(dimethyacetal), which was hydrolyzed to
 8    malonaldehyde and methanol in dilute aqueous solutions in the presence of a strong mineral acid.
 9    In the  drinking water portion of this study, mice were exposed to 3 different concentrations of
10    the malonaldehyde/methanol solution and three different control solutions of methanol alone,
1 1    0.222%, 0.444% and 0.889% methanol in drinking water (222, 444 and 889 ppm, assuming a
12    density of 1  g/ml), corresponding to the stoichiometric amount of methanol liberated by
13    hydrolysis of the acetal in the three test solutions.  The methanol was described as Mallinckrodt
14    analytical grade. No unexposed control groups were included in these studies. However, the
15    author provided pathology data from historical records of untreated Swiss mice of the Eppley
16    colony used in two separate chronic studies, one involving 100 untreated males and 100
17    untreated females (Toth et al., 1977) and the other involving 100 untreated females
18    histopathological analyzed by Apaja (Apaja, 1980).
19          Mice in the Apaja (1980) study were housed five/plastic cage and fed Wayne Lab-Blox
20    pelleted diet. Water was available ad libitum throughout life. Liquid consumption per animal
21    was measured 3 times/week. The methanol dose in the dermal study (females only) was 21.3 mg
22    (532 mg/kg-day using an average weight of 0.04 kg as approximated from Figure 4 of the study),
23    three times/week. The methanol doses in the drinking water study were reported as 22.6, 40.8
                                               4- 1 8       DRAFT-DO NOT CITE OR QUOTE

-------
 1    and 84.5 mg/day (560, 1000 and 2100 mg/kg-day using an average weight of 0.04 kg as
 2    approximated from Figures 14-16 of the study) for females, and 24.6, 43.5 and 82.7 mg/day
 3    (550, 970, and 1800 mg/kg-day using an average weight of 0.045 kg as approximated from
 4    Figures 14-16 of the study) for males,  6 days/week.  The animals were checked daily and body
 5    weights were monitored weekly. The  in-life portion of the experiment ended at 120 weeks with
 6    the death of the last animal.  Like the Soffritti et al. (2002a) study, test animals were sacrificed
 7    and necropsied when moribund.21
 8          The authors reported that survival of the methanol exposed females of the drinking water
 9    study was lower than untreated historical controls  (p < 0.05), but no significant differences in
10    survival was noted for males.  An increase in liver parenchymal cell necrosis was reported in the
11    male and female high-dose groups, with the incidence in females (8%) being significant
12    (p < 0.01) relative to untreated historical controls.  Incidence of acute pancreatitis was higher in
13    high-dose males (p <0.001), but did not appear to be dose-related in females, increasing at the
14    mid- (p <0.0001) and low-doses (p <0.01) when compared to historical controls but not
15    appearing at all in the high-dose females. Significant increases relative to untreated historical
16    controls were noted in amyloidosis of the spleen, nephropathy and pneumonia, but the increases
17    did not appear to be dose related.
18          The author reported incidences of malignant lymphoma in females of 15%, 16%, 36%,
19    and 40% for 532 mg/kg-day (dermal), 560, 1,000,  and 2,100 mg/kg-day (drinking water),
20    respectively. Males from the drinking water study had incidences of malignant lymphoma of 4,
21    24, and 16% for 550, 970, and 1,800 mg/kg-day.  The lymphomas were classified  according to
22    Rappaport's classification (Rappaport,  1966), but location of the lymphoma (organ system) was
23    not reported.  The distributions of lymphomas according to subclasses reported by the author are
24    shown in Table 4-3 for historical untreated and methanol exposed mice in the drinking water
25    studies.  The author indicates that the incidences in both males and females were "within the
26    normal range of occurrence of malignant lymphomas in Eppley Swiss mice," but provides no
27    references or supporting data for this statement and reports elsewhere that the response in high-
28    dose females and mid-dose males were significantly different from unexposed  mice from
29    "historical data of untreated controls (Table 9)" of Toth et al. (1977) (p < 0.05). Though not
30    statistically significant (Fishers exact p = 0.06), the malignant lymphoma response in the mid-
31    dose females was also more than double that of untreated controls from another study (18/100)
32    for which the histopathology was also performed by Apaja (Apaja, 1980).
      21
        The following tisues were fixed in 10% formalin (pH 7.5), embedded in paraffin, sectioned, stained routinely with
      hematoxylineosin (special stains used as needed) and histologically evaluated: skin, lungs, liver spleen, pancreas,
      kidneys, adrenal glands, esophagus, stomach, small and large intestines, rectum, urinary bladder, uterus and ovaries
      or testes, prostate glands and tumors or other gross pathological lesions.
                                                4-19        DRAFT-DO NOT CITE OR QUOTE

-------
          Table 4-3. Incidence of malignant lymphoma responses in Swiss mice exposed to
          methanol in drinking water for life
Malignant Lymphoma
Lymphocytic, well diff.
Lymphocytic moderately diff.
Lymphocytic, poorly diff.
Mixed cell type
Histocytic type
Unclassified
Total
Dose (mg/kg-day)
Males (%)
Oa
n=100






8
550
n=25




4

4
970
n=25

4
4
4
4
8
24C
1800
n=24


8.3

4.2
4.2
17
Females (%)
Oa
n=100






20
Ob
n=100
8
3
7



18
560
n=25


4
4
12

16
1000
n=25
4
4
4
4
12
8
36d
2100
n=25
12
4
4
8
8
4
40C
      aToth et al. (1977);  Hinderer et al. (1979)
      cp < 0.05 as reported by author compared with Toth et al. (1977); Female response is also significant (p < 0.05;
      Fishers exact test) versus untreated controls from Hinderer et al. (1979) and combined controls from both studies.
       p = 0.06 by Fishers exact test versus untreated controls from Hinderer et al. (1979) and combined controls from
      both studies.
      4.2.2. Inhalation Studies
      4.2.2.1. Acute Toxicity
 1          Lewis (1992) reported a 4-hour median lethal concentration (LCso) for methanol in rats of
 2    64,000 ppm (83,867 mg/m3).
 3          Japan's NEDO sponsored a series of toxicological tests on monkeys (M. fascicularis),
 4    rats, and mice, using inhalation exposure.22 A short-term exposure study evaluated monkeys (sex
 5    unspecified) exposed to 3,000 ppm (3,931 mg/m3), 21 hours/day for 20 days (1 animal),
 6    5,000 ppm (6,552 mg/m3) for 5 days (1 animal), 5,000 ppm (6,552 mg/m3) for 14 days (2
 7    animals), and 7,000 and 10,000 ppm (9,173 and 13,104  mg/m3, respectively) for up to 6 days (1
 8    animal at each exposure level) (NEDO,  1987, unpublished report).  Most of the experimental
 9    findings were discussed descriptively in the report, without specifying the extent of change for
10    any of the effects in comparison to seven concurrent controls.  However, the available data
11    indicate that clinical signs of toxicity were apparent in animals exposed to 5,000 ppm (all
12    exposure durations) or higher concentrations of methanol. These included reduced movement,
13    crouching, weak knees, involuntary movements of hands, dyspnea,  and vomiting.  In the
      22 In their bioassays, NEDO (1987) used inbred rats of the F344 or Sprague-Dawley strain, inbred mice of the
      B6C3F1 strain and wild-caught M. fascicularis monkeys imported from Indonesia. The possibility of disease among
      wild-caught animals is a concern, but NEDO (1987) state that the monkeys were initially quarantined for 9 weeks
      and measures were taken throughout the studies against the transmission of pathogens for infectious diseases. The
      authors indicated that "no infectious disease was observed in monkeys" and that "subjects were healthy throughout
      the experiment."
                                                 4-20         DRAFT-DO NOT CITE OR QUOTE

-------
 1    discussion section of the summary report, the authors stated that there was a sharp increase in the
 2    blood levels of methanol and formic acid in monkey exposed to >3,000 ppm (3,931 mg/m3)
 3    methanol.  They reported that methanol and formic acid concentrations in the blood of monkeys
 4    exposed to 3,000 ppm or less were 80 mg/L and 30 mg/L, respectively.23 In contrast, monkeys
 5    exposed to 5,000 ppm or higher concentrations of methanol had blood methanol and formic acid
 6    concentrations of 5,250 mg/L and 1,210 mg/L, respectively. Monkeys exposed to 7,000 ppm and
 7    10,000 ppm became critically ill and had to be sacrificed prematurely. Food intake was said to
 8    be little affected at 3,000 ppm, but those exposed to 5,000 ppm or more showed a marked
 9    reduction.  Clinically, the monkeys exposed to 5,000 ppm or more exhibited reduced movement,
10    weak knees, and involuntary movement of upper extremities, eventually losing consciousness
11    and dying.
12          There were no significant changes in growth, with the exception of animals exposed to
13    the highest concentration, where body weight was reduced by 13%. There were few compound-
14    related changes in hematological or clinical chemistry effects, although animals exposed to 7,000
15    and 10,000 ppm showed an increase in white blood cells.  A marked change in blood pH values
16    at the 7,000 ppm and 10,000 ppm levels (values not reported) was attributed to acidosis due to
17    accumulation of formic acid.  A range of histopathologic changes to the CNS was apparently
18    related to treatment.  Severity of the effects was increased with exposure concentration.  Lesions
19    included characteristic degeneration of the bilateral putamen, caudate nucleus, and claustrum,
20    with associated edema in the cerebral white matter. Necrosis of the basal ganglia was noted
21    following exposure to 5,000 ppm for 5 days (1 animal) and 14 days(l animal).  Exposure to
22    3,000 ppm was considered to be close to the threshold for these necrotic effects, as the monkeys
23    exposed at this level experienced little more than minimal fibrosis of responsive stellate cells of
24    the thalamus, hypothalamus and basal ganglion. The authors reported that no clinical  or
25    histopathological effects of the visual system were apparent, but that exposure to 3,000 ppm
26    (3,931 mg/m3) or more caused dose-dependent fatty degeneration of the liver, and exposure to
27    5,000 ppm (6,552 mg/m3) or more caused vacuolar degeneration of the kidneys, centered on the
28    proximal uniferous tubules.
      4.2.2.2. Subchronic Toxicity
29          A number of experimental studies have examined the effects of subchronic exposure to
30    methanol via inhalation. For example, Sayers et al. (1944) employed a protocol in which 2 male
31    dogs were repeatedly exposed (8 times daily for 3 minutes/exposure) to 10,000 ppm
32    (13,104 mg/m3) methanol for  100 days. One of the dogs was observed for a further 5  days
33    before sacrifice; the other dog was observed for 41  days postexposure.  There were no clinical
      23 Note that Burbacher et al. (2004b, 1999a) measured blood levels of methanol and formic acid in control monkeys
      of 2.4 mg/L and 8.7 mg/L, respectively (see Table 3-3).
                                               4-21        DRAFT-DO NOT CITE OR QUOTE

-------
 1    signs of toxicity, and both gained weight during the study period.  Blood samples were drawn on
 2    a regular basis to monitor hematological parameters, but few if any compound-related changes
 3    were observed. Ophthalmoscopic examination showed no incipient anomalies at any point
 4    during the study period. Median blood concentrations of methanol were 65 mg/L (range 0-280
 5    mg/L) for one dog, and 140 mg/L (70-320 mg/L) for the other.
 6          White et al. (1983) exposed 4 male Sprague-Dawley rats/group, 6 hours/day, 5 days/week
 7    to 0, 200, 2,000, or 10,000 ppm (0, 262, 2,621, and 13,104 mg/m3) methanol for periods of 1, 2,
 8    4, and 6 weeks. Additional groups of 6-week-exposure animals were granted a 6-week
 9    postexposure recovery period prior to sacrifice.  The lungs were excised intact and lavaged
10    6 times with known volumes of physiological saline.  The lavage supernatant was then assayed
11    for lactate dehydrogenase (LDH) and 7V-acetyl-/?-Z)-glucosamidase (/?-NAG) activities.  Other
12    parameters monitored in relation to methanol exposure included absolute and relative lung
13    weights, lung DNA content, protein, acid RNase and acid protease, pulmonary surfactant,
14    number of free cells in lavage/unit lung weight, surface protein, LDH, and /?-NAG. As discussed
15    by the authors, none of the monitored parameters showed significant changes in response to
16    methanol exposure.
17          Andrews et al. (1987) carried out a study of methanol inhalation in 5 Sprague-Dawley
18    rats/sex/group and 3 M. fascicularis monkeys/sex/group, 6 hours/day, 5 days/week, to 0, 500,
19    2,000, or, 5,000 ppm (0, 660, 2,620, and 6,552 mg/m3) methanol for 4 weeks.  Clinical signs
20    were monitored twice daily, and all animals were given a physical examination once a week.
21    Body weights were monitored weekly, and animals received an ophthalmoscopic examination
22    before the start of the experiment and at term. Animals were sacrificed at term by
23    exsanguination following i.v. barbiturate administration.  A gross necropsy was performed,
24    weights of the major organs were recorded, and tissues and organs taken for histopathologic
25    examination. As described by the authors, all animals survived to term with no clinical signs of
26    toxicity among the monkeys and only a few signs of irritation to the eyes and nose among the
27    rats.  In the latter case, instances of mucoid nasal discharges appeared to be dose related. There
28    were no differences in body weight gain among the groups of either rats or monkeys, and overall,
29    absolute and relative organ weights were similar to controls. The  only exception to this was a
30    decrease in the absolute adrenal weight of female high-concentration monkeys and an increase in
31    the relative spleen weight of mid-concentration female rats.  These changes were not considered
32    by the authors to have biological significance. For both rats and monkeys,  there were no
33    compound-related changes in gross pathology, histopathology, or ophthalmoscopy. These data
34    suggest a NOAEL of 5,000 ppm (6,600 mg/m3) for Sprague-Dawley rats and  monkeys under the
35    conditions of the experiment.


                                               4-22        DRAFT-DO NOT CITE OR QUOTE

-------
 1          Two studies by Poon et al. (1995, 1994) examined the effects of methanol on Sprague-
 2    Dawley rats when inhaled for 4 weeks.  The effects of methanol were evaluated in comparison to
 3    those of toluene and toluene/methanol mixtures (Poon et al., 1994), and to gasoline and
 4    gasoline/methanol mixtures (Poon et al., 1995). In the first case (Poon et al., 1994), 10 Sprague-
 5    Dawley rats/sex/group were exposed via inhalation, 6 hours/day, 5 days/week to 0, 300, or
 6    3,000 ppm (0, 393, 3,930 mg/m3) methanol for 4 weeks. Clinical signs were monitored daily,
 7    and food consumption and body weight gain were monitored weekly.  Blood was taken at term
 8    for hematological and clinical chemistry determinations. Weights of the major organs were
 9    recorded at necropsy, and histopathologic examinations were carried out. A 10,000 x g Hver
10    supernatant was prepared from each animal to measure  aniline hydroxylase, aminoantipyrine N-
11    demethylase, and ethoxyresorufin-O-deethylase activities. For the most part, the responses to
12    methanol alone in this experiment were unremarkable.  All animals survived to term, and there
13    were no clinical signs of toxicity among the groups.  Body weight gain and food consumption
14    did not differ from controls,  and there were no compound-related effects in hematological or
15    clinical chemistry parameters or in hepatic mixed function oxidase activities.  However, the
16    authors described a reduction in the size of thyroid follicles that was more obvious in female than
17    male rats.  The authors considered this effect to possibly have been compound related, although
18    the incidence of this feature for the 0, 300, and 3,000 ppm-receiving females was 0/6, 2/6, and
19    2/6, respectively.
20          The second experimental report by Poon et al. (1995) involved the exposure of
21    15 Sprague-Dawley rats/sex/group, 6 hours/day, 5 days/week for 4 weeks to 0 or 2,500 ppm (0
22    and 3,276 mg/m3) to methanol as part of a study on the  toxicological interactions of methanol
23    and gasoline.  Many of the toxicological parameters examined were the same as those described
24    in Poon et  al. (1994) study.  However, in this study urinalysis featured the determination of
25    ascorbic and hippuric  acids.  Additionally, at term, the lungs and tracheae were excised and
26    aspirated with buffer to yield bronchoalveolar lavage fluid that was analyzed for ascorbic acid,
27    protein, and the activities of gamma-glutamyl transferase (y-GT), AP and LDH. Few if any of
28    the monitored parameters showed any differences between controls and those animals exposed to
29    methanol alone. However, two male rats had collapsed right eyes, and there was a reduction in
30    relative spleen weight in females exposed to methanol.  Histopathologic changes in methanol-
31    receiving animals included mild panlobular vacuolation of the liver in females and some mild
32    changes to the upper respiratory tract, including mucous cell metaplasia. The incidence of the
33    latter effect, though higher, was not significantly different than controls in rats exposed to
34    2,500 ppm (3,267 mg/m3) methanol. However, there were also signs of an increased severity of
35    the effect in the presence of the solvent. No histopathologic changes were seen in the lungs or
36    lower respiratory tract of rats exposed to methanol alone.

                                               4-23        DRAFT-DO NOT CITE OR QUOTE

-------
      4.2.2.3. Chronic Toxicity
 1          Information on the chronic toxicity of methanol has come from NEDO (1987,
 2    unpublished report) which includes the results of experiments on 1) monkeys exposed for up to 3
 3    years, 2) rats and mice exposed for 12 months, 3) mice exposed for 18 months, and 4) rats
 4    exposed for 2 years.
 5          In the monkeys, 8 animals (sex unspecified) were exposed to 10, 100, or 1,000 ppm (13,
 6    131, and 1,310 mg/m3) methanol, 21 hours/day, for 7 months (2 animals), 19 months,
 7    (3 animals), or 29 months (3 animals).  There was no indication in the NEDO (1987) report that
 8    this study employed a concurrent control group. One of the 3 animals receiving 100 ppm
 9    methanol and scheduled for sacrifice at 29 months was terminated at 26 months.  Clinical signs
10    were monitored twice daily, body weight changes and food consumption were monitored weekly,
11    and all animals were given a general examination under anesthetic once a month. Blood was
12    collected for hematological  and clinical chemistry tests at term, and all animals were subject to a
13    histopathologic examination of the major organs and tissues.
14          While there were no clinical signs of toxicity in the low-concentration animals, there was
15    some evidence of nasal exudate in monkeys in the mid-concentration group.  High-concentration
16    (1,000 ppm) animals also displayed this response and were observed to scratch themselves over
17    their whole body and crouch for long periods. Food and water intake, body temperature, and
18    body weight changes were the same among the groups. NEDO (1987) reported that there was no
19    abnormality in the retina of any monkey. When animals were examined with an
20    electrocardiogram, there were no abnormalities in the control or 10 ppm groups.  However, in the
21    100 ppm group,  one monkey showed a negative change in the T wave. All 3  monkeys exposed
22    to 1,000 ppm (1,310 mg/m3) displayed this feature, as well as a positive change in the Q wave.
23    This effect was described as a slight myocardial disorder and suggests that 10 ppm (13.1  mg/m3)
24    is a NOAEL for  chronic myocardial effects of methanol and mild respiratory irritation. There
25    were no compound-related effects on hematological parameters. However, 1 monkey in the
26    100 ppm (131 mg/m3) group had greater than normal amounts of total protein, neutral lipids,
27    total and free cholesterol, and glucose, and displayed greater activities of ALT and aspartate
28    transaminase (AST). The authors expressed doubts that these effects were related to methanol
29    exposure and speculated that the animal suffered from liver disease.24
30          Histopathologically, no degeneration of the optical nerve, cerebral cortex, muscles, lungs,
31    trachea, tongue,  alimentary  canal, stomach, small intestine, large intestine, thyroid gland,
32    pancreas, spleen, heart, aorta, urinary bladder, ovary or uterus were reported  (neuropathological
33    findings are discussed below Section 4.4.2. Most of the internal organs showed no compound-
      24 Ordinarily, the potential for liver disease in test animals would be remote, but may be a possibility in this case
      given that these monkeys were captured in the wild.
                                               4-24        DRAFT-DO NOT CITE OR QUOTE

-------
 1    related histopathologic lesions.  However, there were signs of incipient fibrosis and round cell
 2    infiltration of the liver in monkeys exposed to 1,000 ppm (1,310 mg/m3) for 29 months. NEDO
 3    (1987) indicated that this fibrosis occurred in 2/3 monkeys of the 1,000 ppm group to a "strictly
 4    limited extent."  They also qualitatively reported a dose-dependent increase in "fat granules" in
 5    liver cells "centered mainly around the central veins" at all doses, but did not  provide any
 6    response data. The authors state that 1,000 ppm (1,310 mg/m3) represents a chronic lowest-
 7    observed-adverse-effect level (LOAEL) for hepatic effects of inhaled methanol, suggesting that
 8    the no effect level would be 100 ppm (131 mg/m3).  However, this is a tenuous determination
 9    given the lack of information on the pathological progression and significance of the appearance
10    of liver cell fat granules at exposures below 1,000 ppm and the lack detail (e.g., time of sacrifice)
11    for the control group.
12          Dose-dependent changes were observed in the kidney; NEDO (1987) described the
13    appearance of Sudan-positive granules in the renal tubular epithelium at 100 ppm (131 mg/m3)
14    and 1,000 (1,310 mg/m3) and hyalinization of the glomerulus and penetration of round cells into
15    the renal tubule stroma of monkeys exposed to methanol at 1,000 (1,310 mg/m3).  The former
16    effect was more marked at the higher concentration and was thought by the authors to be
17    compound-related. This would indicate a no  effect level at 10 ppm (13.1 mg/m3) for the chronic
18    renal effects of methanol.  The authors observed atrophy of the tracheal epithelium in four
19    monkeys. However, the incidence of these effects was unrelated to dose and therefore, could not
20    be unequivocally ascribed to an effect of the solvent. No other histopathologic abnormalities
21    were related to the effects of methanol.  Confidence in these determinations is considerably
22    weakened by concern over whether a concurrent control group was used in the chronic study.25
23          NEDO (1987) describes a 12-month inhalation study  in which 20 F344 rats/sex/group
24    were exposed to 0, 10, 100, or 1,000 ppm (0,  13.1, 131, and 1,310 mg/m3) methanol,
25    approximately 20 hours/day, for a year.  Clinical signs of toxicity were monitored daily; body
26    weights and food consumption were recorded weekly for the first 13 weeks, then monthly.
27    Blood samples were drawn at term to measure hematological and clinical chemistry parameters.
28    Weights of the major organs were monitored  at term, and a histopathologic examination was
29    carried out on all major organs and tissues. Survival was high among the groups;  one high-
30    concentration female died  on day 337 and one low-concentration male died on day 340. As
31    described by the authors, a number of procedural anomalies arose during this  study.  For
32    example, male controls in two cages lost weight because of an interruption to the water supply.
33    Another problem was that the brand of feed was changed during the study. Fluctuations in some
34    clinical chemistry and hematological parameters were recorded.  The authors considered the
35    fluctuations to be minor and within the normal range.  Likewise, a number of histopathologic
      25 All control group responses were reported in a single table in the section of the NEDO (1987) report that describes
      the acute monkey study, with no indication as to when the control group was sacrificed.
                                               4-25        DRAFT-DO NOT CITE OR QUOTE

-------
 1    changes were observed, which, in every case, were considered to be unrelated to exposure level
 2    or due to aging.
 3           A companion experiment featured the exposure of 30 B6C3F1 mice/sex/group for 1 year
 4    to the same concentrations as the F344 rats (NEDO, 1987).  Broadly speaking, the same suite of
 5    toxicological parameters was monitored as described above, with the addition of urinalysis.
 6    10 mice/sex/group were sacrificed at 6 months to provide interim data on the parameters under
 7    investigation. A slight atrophy in the external lacrimal gland was observed in both sexes and was
 8    significant in the 1,000 ppm male group compared with controls. An apparently dose-related
 9    increase in moderate fatty degeneration of hepatocytes was observed in males (1/20, 4/20, 6/20
10    and 8/20 in the 0, 10, 100, and 1,000 ppm dose groups, respectively) which was significantly
11    increased over controls at the  1,000 ppm dose. However the incidence of moderate to severe
12    fatty degeneration was observed in untreated animals maintained outside of the chamber. In
13    addition, there was a clear correlation between fatty degeneration and body weight (a change
14    which was not associated with treatment at 12 months); heavier animals tended to have more
15    severe cases of fatty degeneration. The possibility of renal deficits due to methanol exposure
16    was suggested by the appearance of protein in the urine.  However, this effect was also seen in
17    controls and did not display a  dose-response effect.  Therefore, it is unlikely to be a consequence
18    of exposure to methanol. NEDO (1987) reported other histopathologic and biochemical (e.g.,
19    urinalysis and hematology) findings that do not appear to be related to treatment, including a
20    number of what were considered to be spontaneous tumors in both control and exposure groups.
21           NEDO (1987, 1985/2008a, unpublished reports)26 exposed 52 male and 53 female
22    B6C3F1 mice/group for 18 months at the same concentrations of methanol (0, 10, 100 and
23    1,000 ppm) and with a similar experimental protocol to that described in the 12-month studies.27
24    The fact that the duration of this study was only 18 months and not the more typical 2 years
25    limits its ability to detect carcinogenic responses with relatively long latency periods.  Animals
26    were sacrificed at the end of the 18-month exposure period. NEDO (1985/2008a) reported that
27    "there was no microbiological contamination that may have influenced the result of the study"
28    and that the study included an assessment of general conditions, body weight change, food
29    consumption rate, laboratory tests (urinalysis, hematological, and plasma biochemistry) and
30    pathological tests (pathological autopsy,28 organ weight check and histopathology29). As stated
      26 This study is described in a summary report (NEDO, 1987) and a more detailed, eight volume translation of the
      original chronic mouse study report (NEDO, 1985/2008a).
      27 The authors reported that "[t]he levels of methanol turned out to be ~4 ppm in low level exposure group (10 ppm)
      for ~11 weeks from week 43 of exposure due to the analyzer malfunction" and that "the average duration of
      methanol exposure was 19.1 hours/day for both male and female mice."
      28 Autopsy was performed on all cases to look for gross lesions in each organ.
                                                4-26        DRAFT-DO NOT CITE OR QUOTE

-------
 1    in the summary report (NEDO, 1987), a few animals showed clinical signs of toxicity, but the
 2    incidence of these responses was not related to dose. Likewise, there were no compound-related
 3    changes in body weight increase, food consumption,30 urinalysis, hematology,  or clinical
 4    chemistry parameters. High-concentration males had lower testis weights compared to control
 5    males. Significant differences were detected for both absolute and relative testis weights. One
 6    animal in the high-dose group had severely atrophied testis weights, approximately 25% of that
 7    of the others in the dose group. Exclusion of this animal in the analysis still resulted in a
 8    significant difference in absolute testis weight compared to controls but resulted in no difference
 9    in relative testis weight. High-concentration females had higher absolute kidney and spleen
10    weights compared to controls, but there was no significant difference in these organ weights
11    relative to body weight. At necropsy, there were signs  of swelling in spleen, preputial glands,
12    and uterus in some animals. Some  animals developed nodes in the liver and lung although,
13    according to the authors, none of these changes were treatment-related. NEDO (1985/2008a)
14    reported that all nonneoplastic changes were "nonspecific and naturally occurring changes that
15    are often experienced by 18-month old B6C3F1 mice"  and that fatty degeneration of liver that
16    was suspected to occur dose-dependently in the 12-month NEDO (1987) study was not observed
17    in this study.  Similarly, though the study found various neoplastic changes across dose groups,
18    there was no compound-related formation of tumors in any organ or tissue.
19           EPA reviewed the cancer findings documented in a recent translation of the original
20    NEDO report on this chronic mouse study (NEDO, 1985/2008a) to identify possible compound-
21    related effects.  Hyperplastic and neoplastic histopathological findings have been tabulated and
22    are as shown in Table 4-4.
      29 Complete histopathological examinations were performed for the control group and high-dose (1,000 ppm)
      groups. Only histopathological examinations of the liver were performed on the low- and medium-level exposure
      groups because no chemical-related changes were found in the high-level exposure group and because liver changes
      were noted in the 12-month mouse study (NEDO, 1987).
      30 NEDO (1985/2008a) reports sporadic reductions in food consumption of the 1,000 ppm group, but no associated
      weight loss or abnormal test results.
                                                 4-27         DRAFT-DO NOT CITE OR QUOTE

-------
          Table 4-4. Histopathological changes in tissues of B6C3F1 mice exposed to methanol
          via inhalation for 18 months
Tissues/tumor type
Exposure concentration (ppm)
0
10
100
1,000
0
10
100
1,000
Number of animals affected/number examined
Males
Females
Lung
Adenomatosis
Pulmonary adenoma
0/52
4/52
0/3
0/3
0/3
0/3
0/52
7/52
0/53
3/53
0/0
0/0
0/5
0/0
1/53
2/53
Liver
Hepatocellular adenoma
Hepatocellular carcinoma
Neoplastic nodule
3/52
2/52
16/52
2/52
4/52
13/52
2/52
0/52
16/52
4/52
1/52
20/52
1/53
3/53
1/53
1/52
0/52
0/52
1/53
3/53
0/53
4/53
2/53
1/53
      Source: NEDO (1985/2008a).

 1          There is no clear evidence for treatment-related carcinogenic effects in the mice in this
 2    study. However, the fact that the study duration was limited to 18 months rather than the
 3    traditional 2-year bioassay makes it difficult to draw a definitive conclusion, particularly
 4    regarding pulmonary adenomas, which were marginally increased in high-dose male mice of this
 5    study and were also increased in male rats of the NEDO chronic rat study (NEDO, 1987,
 6    1985/2008b, unpublished reports). In this study, the lack of adenomatosis in control or treated
 7    male mice supports the conclusion of the authors that the observed tumors were probably
 8    unrelated to methanol exposure.  There was no apparent relationship to treatment in any
 9    neoplastic findings in the liver. Of relevance to the findings of treatment-related lymphomas and
10    leukemias in Sprague-Dawley rats receiving methanol in drinking water in the Soffritti et al.
11    (2002a) study, few lymphomas and leukemias were identified in the NEDO (1987, 1985/2008b)
12    study reports, with no sign of a dose-related trend.
13          Another study reported in (NEDO (1987, 1985/2008b)31 was a 24-month carcinogenicity
14    bioassay in which 52 F344 rats/sex/group were kept in whole body inhalation chambers
15    containing 0, 10,  100, or 1,000 ppm (0, 13.1, 131, and 1,310 mg/m3) methanol vapor. Animals
16    were maintained in the exposure chambers for approximately 19.5 hours/day for a total of 733-
17    736 days (males) and 740-743 days (females). Animals were monitored once a day for clinical
18    signs of toxicity, body weights were recorded once a week, and food consumption was measured
19    weekly in a 24-animal subset from each group.  Urinalysis was carried out on the day prior to
20    sacrifice for each animal, the samples being monitored for pH, protein, glucose, ketones,
21    bilirubin, occult blood, and urobilinogen. Routine clinical chemistry and hematological
      31 This study is described in a summary report (NEDO, 1987) and a more detailed, eight-volume translation of the
      original chronic rat study report (NEDO, 1985/2008b).
                                               4-28        DRAFT-DO NOT CITE OR QUOTE

-------
 1    measurements were carried out and all animals were subject to necropsy at term, with a
 2    comprehensive histopathological examination of tissues and organs.32
 3           There was some fluctuation in survival rates among the groups in the rat study, though
 4    apparently unrelated to exposure concentration.33 In all groups, at least 60% of the animals
 5    survived to term. A number of toxicological responses were described by the authors, including
 6    atrophy of the testis, cataract formation, exophthalmia, small eye ball, alopecia, and paralysis of
 7    the hind leg.  However, according to the authors, the incidence of these effects were unrelated to
 8    dose and more likely represented effects of aging. NEDO (1985/2000b) reported a mild,
 9    nonsignificant (4%) body weight suppression among 1,000 ppm females between 51 and
10    72 weeks, but that body weight gain was largely similar among the groups for the duration of the
11    experiment. Food consumption was significantly lower than controls in high-concentration male
12    rats during the day 210-365 time interval, but no corresponding weight loss was observed.
13    Among hematological parameters, mid- and high-concentration females had a significantly
14    (p> 0.05) higher differential leukocyte count than controls, but dose dependency was not
15    observed. Serum total  cholesterol, triglyceride, free fatty acid, and phospholipid concentrations
16    were significantly (p > 0.05) lower in high-concentration females compared to controls.
17    Likewise, serum sodium concentrations were significantly (p > 0.05) lower in mid- and high-
18    concentration males compared to controls.  High-concentration females had significantly lower
19    (p > 0.05) serum concentrations of inorganic phosphorus but significantly (p > 0.05) higher
20    concentrations of potassium compared to controls. Glucose levels were elevated in the urine of
21    high-concentration male rats relative to controls, and female rats had lower pH values and higher
22    bilirubin levels in mid- and high-concentration groups relative to controls. In general, NEDO
23    (1987,  1985/2008b) reported that these variations in urinary, hematology, and clinical chemistry
24    parameters were not related to chemical exposure.
25           NEDO (1987)  reported that there was little change in absolute or relative weights of the
26    major organs or tissues.  When the animals were examined grossly at necropsy, there were some
27    signs of swelling in  the pituitary and thyroid, but these effects were judged to be unrelated to
28    treatment. The most predominant effect was the dose-dependent formation of nodes in the lung
29    of males (2/52, 4/52, 5/52, and 10/52 \p < 0.01] for control, low-, mid-, and high-concentration
30    groups, respectively).  Histopathologic examination pointed to a possible association of these
31    nodes with the appearance of pulmonary adenoma (1/52, 5/52, 2/52, and 6/52 for control, low-,
32    mid- and high-concentration groups, respectively) and a single pulmonary adenocarcinoma in the
      32 Complete histopathological examinations were performed on the cases killed on schedule (week 104) among the
      control and high-exposure groups, and the cases that were found dead/ killed in extremis of all the groups. Because
      effects were observed in male and female kidneys, male lungs as well as female adrenal glands of the high-level
      exposure group, these organs were histopathologically examined in the low- and mid-exposure groups.
      33Survival at the time of exposure termination (24 months) was 69%, 65%, 81%, and 65% for males and 60%, 63%,
      60% and 67% for females of the control, low-, mid- and high-exposure groups, respectively.
                                                4-29         DRAFT-DO NOT CITE OR QUOTE

-------
 1    high-dose group (1/52).  Other examples of tumor formation that were increased in high-
 2    concentration animals versus controls included an increased incidence of pituitary adenomas in
 3    high-concentration males (17/52 compared to 12/52 controls), hyperplastic change in the testis in
 4    high-concentration males (10/52 compared to 4/52 controls), and chromaffmoma
 5    (pheochromocytomas)34 in the adrenals of high-concentration females (7/52 compared to 2/52
 6    controls). Individually, these changes did not achieve statistical significance, and in general, the
 7    authors concluded that few if any of the observed changes were effects of methanol.
 8          EPA reviewed the cancer findings of this study that are documented in a recent translation
 9    of the original NEDO (1985/2008b) report to identify possible compound-related effects. High-
10    dose incidences of pituitary adenomas (17/52;  33%) and hyperplastic change in testes (10/52;
11    19%) mentioned above were within historical incidences for this rat strain.35  However, the
12    observed incidence rate for pulmonary adenoma/adenocarcinoma in high-dose males of 13.5%
13    (7/52) was significantly elevated (Fisher's exact test/? < 0.05) over the concurrent control rate of
14    2% (1/52) and historical control rates of 2.5% ± 2.6% (n = 1054) and 3.84% ± 2.94% (n = 1199)
15    reported by NTP for the pre-1995 control F344 male rats fed NIH-07 diet (NTP, 1999) and  post-
16    1994 control F344 male rats fed NTP-2000 diet (NTP, 2007), respectively. Also, the incidence of
17    pulmonary adenoma/adenocarcinoma in male rats exhibited a dose-response trend (Cochrane-
18    Armitage/? < 0.05). While the observed incidence rate for pheochromocytomas in high-dose
19    females of 13.7% (7/51) was not significantly elevated over the concurrent control rate of 4%
20    (2/50), it was significantly elevated (Fisher's exact test/? < 0.05) over NTP historical control
21    rates for total (benign, complex and malignant) pheochromocytomas of 2.5% ± 2.6% (n = 1054)
22    and 3.84% ± 2.94% (n = 1199) reported by NTP for pre-1995 control F344 female rats fed NIH-
23    07 diet (NTP, 1999) and post-1994 control F344 female rats fed NTP-2000 diet (NTP, 2007),
24    respectively.36  Also, the incidence of pheochromocytomas in female rats exhibited a dose-
25    response trend (Cochrane-Armitage/? < 0.05).  The histopathological incidences for pulmonary
26    and adrenal effects reported by NEDO (1987, 1985/2008b) are shown in Table 4-5.
      34 There were some differences in nomenclature used in the NEDO (1985/2008b) report translation versus those
      used in the older summary report (NEDO, 1987). For example, it is probable that the adrenal chromaffinoma
      referred to in NEDO (1987) are the same lesions as the pheochromocytoma referred to in NEDO (1985/2008b).
      35 NTP reports high incidences in historical control male F344 rats of pituitary gland adenomas, ranging from 45.4%
      ± 20.19% (NTP, 2007) to 63.4% ± 18.3% (NTP, 1999). While control incidences for testicular hyperplasia are not
      reported, historical incidences of testicular ademoma ranged from 70.1% ±11.2% (NTP, 1999) to 86.32% ±9.34%
      (NTP, 2007) in this rat strain.
      36 NEDO (1987, 1985/2008b) does not categorize reported chromoffinoma (pheochromocytomas) as benign,
      complex or malignant. The historical rates for complex and malignant tumors are much lower, ranging from 0.1%
      to 0.7 % for female F344 rats (NTP, 2007; NTP, 1999; Haseman et al., 1998).
                                                4-30        DRAFT-DO NOT CITE OR QUOTE

-------
          Table 4-5. Histopathological changes in lung and adrenal tissues of F344 rats exposed
          to methanol via inhalation for 24 months
Tissues/
tumor type
Exposure concentration (ppm)
0
10
100
1000
0
10
100
1000
Number of animals affected/number examined
Males
Females
Lung
Pulmonary adenoma
Pulmonary adenocarcinoma
Combined pulmonary
adenoma/adenocarcinoma
Adenomatosis
Epithelial swelling
1/52
0/52
1/52
4/52
3/52
5/50
0/50
5/50
1/50
2/50
2/52
0/52
2/52
5/52
1/52
6/52
1/52
7/52 a'b
4/52
1/52
2/52
0/52
2/52
3/52
0/52
0/19
0/19
0/19
2/19
0/19
0/20
0/20
0/20
1/20
0/20
0/52
0/52
0/52
1/52
0/52
Adrenal glands
Pheochromocytoma
Medullary hyperplasia
7/52
0/52
2/16
0/16
2/10
0/10
4/51
2/51
2/50
2/50
3/51
3/51
2/49
7/49
7/51b'c
2/51
     ap < 0.05 over concurrent controls using the Fisher's Exact test.
     bp < 0.05 for Cochrane-Armitage test of overall dose-response trend.
     °p < 0.05 over NTP historical controls for total (benign, complex and malignant) pheochromocytomas using the
     Fisher's Exact test

     Source: NEDO (1987, 1985/2008b).

 1          In contrast to the conclusions of the NEDO (1987) summary report that there were no
 2   compound-related changes in F344 rats exposed to methanol via inhalation, EPA identifies
 3   potential treatment-related changes in the lungs of male rats and the adrenal medulla of female
 4   rats in the more detailed translation of the original report (NEDO,  1985/2008b). The NEDO
 5   (1987) summary report did not report the statistically significant combined pulmonary adenoma
 6   and adenocarcinoma finding in the high-dose group of male rats. Table 6 (page 146) of the
 7   NEDO (1987) summary reports only "Tumural changes occurring at a rate of over 5%." The
 8   lung response of the male rats as  shown in Table 4-5 suggests a proliferative change in cells of
 9   the alveolar epithelium involving a progression towards adenoma and adenocarcinoma that
10   appears to be more pronounced with increasing methanol exposure and considerably elevated
11   over historical controls.  Similarly, for female rats, the observed increase in medullary
12   hyperplasia in the 100 ppm dose  group, in conjunction with a higher incidence of
13   pheochromocytoma in the adrenal gland is suggestive of a methanol-induced progressive change
14   leading to a carcinogenic response.
                                               4-31
DRAFT-DO NOT CITE OR QUOTE

-------
      4.3. REPRODUCTIVE AND DEVELOPMENTAL STUDIES-ORAL AND INHALATION
 1         Many studies have been conducted to investigate the reproductive and developmental
 2    toxicity of methanol.  The purpose of these studies was principally to determine if methanol has
 3    a similar toxicology profile to another widely studied teratogen, ethanol.

      4.3.1. Oral Studies
 4          Three studies were identified that investigated the reproductive and developmental effects
 5    of methanol in rodents via the oral route (Fu et al., 1996; Sakanashi et al., 1996; Rogers et al.,
 6    1993a). Two of these studies also investigated the influence of folic acid-deficient (FAD) diets on
 7    the effects of methanol exposures (Fu et al., 1996; Sakanashi et al., 1996).
 8          Rogers et al. (1993a) conducted a developmental toxicity study in which methanol  in
 9    water was administered to pregnant female CD-I mice via gavage on GD6-GD15. Eight test
10    animals received 4 g/kg-day  methanol given in 2 daily doses of 2g/kg; 4 controls received
11    distilled water. By analogy to the protocol of an inhalation study of methanol that was described
12    in the same report, it is assumed that dams were sacrificed on GD17,  at which point implantation
13    sites, live and dead fetuses, resorptions/litter, and the incidences of external and skeletal
14    anomalies and malformations were determined. In the brief summary of the findings provided
15    by the authors, it appears that cleft palate (43.5% per litter versus 0% in controls) and
16    exencephaly (29% per litter versus 0% in controls) were the prominent external defects
17    following maternal methanol exposure by gavage. Likewise, an increase in totally resorbed
18    litters and a decrease in the number of live fetuses per litter were evident. However, it is possible
19    that these effects may have been caused or exacerbated by the high bolus dosing regimen
20    employed. It is also possible that effects were not observed due to the limited study size. The
21    small number of animals in the control group relative to the test group limits the power of this
22    study to detect treatment-related responses.
23          Sakanashi et al. (1996) tested the influence of dietary folic acid intake on various
24    reproductive and developmental effects observed in CD-I mice exposed to methanol.  Starting
25    5 weeks prior to breeding and continuing for the remainder of the study, female CD-I mice were
26    fed folic acid free diets supplemented with 400 (low), 600 (marginal) or 1,200 (sufficient) nmol
27    folic acid/kg. After 5 weeks on their respective diets, females were bred with CD-I male mice.
28    On GD6-GD15, pregnant mice in each of the diet groups were given twice-daily gavage doses of
29    2.0 or 2.5 g/kg-day methanol (total dosage of 4.0 or 5.0 g/kg-day).  On GDIS, mice were
30    weighed and killed, and the liver, kidneys and gravid uteri removed and weighed. Maternal liver
31    and plasma folate levels were measured, and implantation sites, live and dead fetuses, and
32    resorptions were counted. Fetuses were weighed individually and examined for cleft palate and
33    exencephaly. One third of the fetuses in each litter were examined for skeletal morphology.
                                               4-32       DRAFT-DO NOT CITE OR QUOTE

-------
 1    They observed an approximate 50% reduction in liver and plasma folate levels in the mice fed
 2    low versus sufficient folic acid diets in both the methanol exposed and unexposed groups.
 3    Similar to Rogers et al. (1993a), Sakanashi et al. (1996) observed that an oral dose of 4-5 g/kg-
 4    day methanol during GD6-GD15 resulted in an increase in cleft palate in mice fed sufficient
 5    folic acid diets, as well as an increase in resorptions and a decrease in live fetuses per litter. They
 6    did not observe an increase in exencephaly in the FAS group at these doses, and the authors
 7    suggest that this may be due to diet and the source of CD-I mice differing between the two
 8    studies.
 9          In the case of the animals fed the folate deficient diet, there was a 50% reduction in
10    maternal liver folate concentration and a threefold increase in the percentage of litters affected by
11    cleft palate (86.2% versus 34.5% in mice fed sufficient folic acid) and a 10-fold increase in the
12    percentage of litters affected by exencephaly (34.5% versus 3.4% in mice fed sufficient folic
13    acid) at the 5 g/kg methanol dose. Sakanashi et al. (1996) speculate that the increased methanol
14    effect from the FAD diet could have been due to an increase in tissue formate levels (not
15    measured) or to a critical reduction in conceptus folate concentration following the methanol
16    exposure. Plasma and liver folate levels at GDIS within each dietary group were not
17    significantly different between exposed versus unexposed mice. However, these measurements
18    were taken 3 days after methanol exposure. Dorman et al. (1995) observed a transient decrease
19    in maternal red blood cells (RBCs) and conceptus folate levels within 2 hours following
20    inhalation exposure to 15,000 ppm methanol on GD8. Thus, it is possible that short-term
21    reductions in available folate during GD6-GD15 may have affected fetal development.
22          Fu et al. (1996) also tested the influence of dietary folic acid intake on reproductive and
23    developmental effects observed in CD-I mice exposed to methanol. This study was performed
24    by the same laboratory and used a  similar study design and dosing regimen as Sakanashi et al.
25    (1996), but exposed the pregnant mice to only the higher 2.5 g/kg-day methanol (total dosage of
26    5.0 g/kg-day) on GD6-GD10.  Like Sakanashi et al. (1996), Fu et al. (1996) measured maternal
27    liver and plasma folate levels on GDIS and observed similar, significant reductions in these
28    levels for the FAD versus FAS mice. However, Fu et al.  (1996) also measured fetal liver folate
29    levels at GDIS.  This measurement does not address the question of whether methanol exposure
30    caused short-term reductions in fetal liver folate because it was taken 8 days after the GD6-
31    GD10 exposure period.  However, it did provide evidence regarding the extent to which a
32    maternal FAD diet can impact fetal liver folate levels in this species and strain. Significantly, the
33    maternal FAD diet had a greater impact on fetal liver folate than maternal liver folate levels.
34    Relative to the FAS groups, fetal liver folate levels in the FAD groups were reduced 2.7-fold for
35    mice not exposed to methanol (1.86 ± 0.15 nmol/g in the FAD group versus 5.04 ± 0.22 nmol/g
36    in the FAS group) and 3.5-fold for mice exposed to methanol (1.69 ± 0.12 nmol/g in the FAD

                                               4-33        DRAFT-DO NOT CITE OR QUOTE

-------
 1    group versus 5.89 ± 0.39 nmol/g in the FAS group).  Maternal folate levels in the FAD groups
 2    were only reduced twofold both for mice not exposed (4.65 ± 0.37 versus 9.54 ± 0.50 nmol/g)
 3    and exposed (4.55 ± 0.19 versus 9.26 ± 0.42 nmol/g). Another key finding of the Fu et al. (1996)
 4    study is that methanol exposure during GD6-GD10 appeared to have similar fetotoxic effects,
 5    including cleft palate, exencephaly, resorptions, and decrease in live fetuses, as the same level of
 6    methanol exposure administered during GD6-GD15 (Sakanashi et al., 1996; Rogers et al.,
 7    1993a).  This is consistent with the hypothesis made by Rogers et al. (1993b) that the critical
 8    period for methanol-induced cleft palate and exencephaly in CD-I mice is within GD6-GD10.
 9    As in the studies of Sakanashi et al. (1996) and Rogers et al. (1993a), Fu et al. (1996) reported a
10    higher incidence of cleft palate than exencephaly.

      4.3.2. Inhalation Studies
11          Nelson et al. (1985) exposed 15 pregnant Sprague-Dawley rats/group to 0, 5,000, 10,000,
12    or 20,000 ppm (0, 6,552, 13,104, and 26,209 mg/m3) methanol (99.1% purity) for 7 hours/day.
13    Exposures were conducted on GDI-GDI9 in the two lower concentration groups and GD7-
14    GDIS in the highest concentration group, apparently on separate days. Two groups of 15 control
15    rats were exposed to air only.  Day 1 blood methanol levels measured 5 minutes after the
16    termination of exposure in NP rats that had received the same concentrations of methanol as
17    those animals in the main part of the experiment were 1.00 ± 0.21, 2.24 ± 0.20, and 8.65 ± 0.40
18    mg/mL for those exposed to 5,000, 10,000 and 20,000 ppm methanol, respectively. Evidence of
19    maternal toxicity included a slightly unsteady gait in the 20,000 ppm group during the first few
20    days  of exposure. Maternal body weight gain and food intake were unaffected by methanol.
21    Dams were sacrificed on GD20, and 13-30  litters/group were evaluated.  No effect was observed
22    on the number of corpora lutea or implantations or the percentage of dead or resorbed fetuses.
23    Statistical evaluations included analysis of variance (ANOVA) for body weight effect, Kruskal-
24    Wallis test for endpoints such as litter size and viability and Fisher's exact test for
25    malformations. Fetal body weight was significantly reduced at concentrations of 10,000 and
26    20,000 ppm by 7% and 12-16%, respectively, compared to controls. An increased number of
27    litters with skeletal and visceral malformations were observed at > 10,000 ppm, with statistical
28    significance obtained at 20,000 ppm.  Numbers of litters with visceral malformations were 0/15,
29    5/15, and 10/15 and with skeletal malformations were 0/15, 2/15, and 14/15 at 0, 10,000, and
30    20,000 ppm, respectively. Visceral malformations included exencephaly and encephaloceles.
31    The most frequently observed skeletal malformations were rudimentary and extra cervical ribs.
32    The developmental and maternal NOAELs for this study were identified  as 5,000 ppm (6,552
33    mg/m3) and 10,000 ppm (13,104 mg/m3), respectively.

                                              4-34         DRAFT-DO  NOT CITE OR QUOTE

-------
 1          NEDO (1987) sponsored a teratology study in Sprague-Dawley rats that included an
 2    evaluation of postnatal effects in addition to standard prenatal endpoints. Thirty-six pregnant
 3    females/group were exposed to 0, 200, 1,000, or 5,000 ppm (0, 262, 1,310, and 6,552 mg/m3)
 4    methanol vapors (reagent grade) on GD7-GD17 for 22.7 hours/day. Statistical significance of
 5    results was evaluated by t-test, Mann-Whitney U test, Fisher's exact test, and/or Armitage's $
 6    test.
 7          Contrary to the Nelson et al. (1985) report of a 10,000 ppm NOAEL for this rat strain, in
 8    the prenatal portion of the NEDO (1987) study, reduced body weight gain and food and water
 9    intake during the first 7 days of exposure were reported for dams in the 5,000 ppm group.
10    However, it was not specified if these results were statistically significant. One dam in the
11    5,000 ppm group died on GDI9, and  one dam was sacrificed on GDIS in moribund condition.
12    On GD20, 19-24 dams/group were sacrificed to evaluate the incidence of reproductive deficits
13    and such developmental parameters as fetal viability, weight, sex, and the occurrence of
14    malformations. As summarized in Table 4-6, adverse reproductive and fetal effects were limited
15    to the 5,000 ppm group and included  an increase in late-term resorptions, decreased live fetuses,
16    reduced  fetal weight, and increased frequency of litters with fetal malformations, variations, and
17    delayed  ossifications. Malformations or variations included defects in ventricular septum,
18    thymus,  vertebrae, and ribs.
19          Postnatal effects of methanol inhalation were evaluated in the remaining 12 dams/group
20    that were permitted to deliver and nurse their litters. Effects were only observed in the
21    5,000 ppm group, and included a 1-day prolongation of the gestation period and reduced post-
22    implantation survival, number of live pups/litter, and survival on PND4 (Table 4-7). When the
23    delay in  parturition was considered, methanol treatment had no effect on attainment of
24    developmental milestones such as eyelid opening, auricle development, incisor eruption, testes
25    descent,  or vaginal opening.  There were no adverse body weight effects in offspring from
26    methanol treated groups. The weights of some  organs (brain, thyroid, thymus, and testes) were
27    reduced  in 8-week-old offspring exposed to 5000 ppm methanol during prenatal development.
28
29
30
31
                                               4-3 5        DRAFT-DO NOT CITE OR QUOTE

-------
1
2
3
4
         Table 4-6. Reproductive and developmental toxicity in pregnant Sprague-Dawley rats
         exposed to methanol via inhalation during gestation
Effect
Exposure concentration (ppm)
0
200
1,000
5,000
Reproductive effects
Number of pregnant females examined
Number of corpora lutea
Number of implantations
Number of resorptions
Number of live fetuses
Sex ratio (M/F)
Fetal weight (male)
Fetal weight (female)
Total resorption rate (%)
19
17.0 ±2.6
15.7±1.6
0.79 ±0.85
14. 95 ±1.61
144/140
3.70 ±0.24
3. 51 ±0.19
11.2±9.0
24
17.2 ±2.7
15.0 ±3.0
0.71 ±1.23
14.25 ±3. 54
177/165
3.88 ±0.23
3.60 ±0.25
15.6±21.3
22
16.4 ±1.9
15.5 ±1.2
0.95 ±0.65
14. 55 ±1.1
164/156
3.82 ±0.29
3.60 ±0.30
10.6 ±8.4
20
16. 5 ±2.4
14.5 ±3.3
1.67 ±2.03
12.86±4.04a
134/136
3.02±0.27C
2.83±0.26C
23.3±22.7a
Soft tissue malformations
Number of fetuses examined
Abnormality at base of right subclavian
Excessive left subclavian
Ventricular septal defect
Residual thymus
Serpengious urinary tract
136
0.7 ±2.87(1)
0
0
2.9 ±5. 91 (4)
43.0 ±24.64 (18)
165
0
0
0.6 ±2. 96(1)
2.4 ±5.44 (4)
35.2 ±31.62 (19)
154
0
0
0
2.6 ±5.73 (4)
41. 8 ±38.45 (15)
131
0
3.5 ±9.08 (3)
47.6 ± 36.51 (16)b
53.3±28.6(20)b
22.1 ±22.91 (13)
Skeletal abnormalities
Number of fetuses examined
Atresia of foramen costotransversarium
Patency of foramen costotransversium
Cleft sternum
Split sternum
Bifurcated vertebral center
Cervical rib
Excessive sublingual neuropore
Curved scapula
Waved rib
Abnormal formation of lumbar
vertebrae
148
23. 5 ±5.47 (3)
0
0
0
0.8 ±3.28(1)
0
0
0
0
0
177
7.7 ±1.3 (8)
0
0
0
1.6 ±5.61 (2)
0
0
0
0
0
165
3. 5 ±8. 88 (4)
0.6 ±2.67(1)
0
0
3.0 ±8.16 (3)
0
0
0
0
0
138
45.2±25.18(20)b
13.7 ±20.58 (7)
5.6 ±14. 14 (3)
7.0 ±14.01 (5)
14.5 ± 16.69 (ll)b
65.2±25.95(19)b
49. 9 ±27.31 (19)
0.7±3.19(1)
6.1 ±11. 84 (5)
0.7±3.19(1)
      ap < 0.05, bp < 0.01, °p < 0.001, as calculated by the authors.
      Values are means ± S.D. Values in parentheses are the numbers of litters.
      Source: NEDO (1987).
                                               4-36
DRAFT-DO NOT CITE OR QUOTE

-------
          Table 4-7. Reproductive parameters in Sprague-Dawley dams exposed to methanol
          during pregnancy then allowed to deliver their pups
Effect
Number of dams
Duration of gestation (days)
Number of implantations
Number of pups
Number of live pups
Number of live pups on PND4
Sex ratio (M/F)
Postimplantation embryo
survival rate
Exposure concentration (ppm)
0
12
21.9 ±0.3
15.8 ±1.6
15.2 ±1.6
15.2 ±1.6
15.0 ±1.7 (2)
88/94
96.3 ±4.2
200
12
21.9 ±0.3
14.8 ±1.2
14.4 ±1.3
14.1 ±1.4
13.8 ±1.5 (3)
87/85
94.9 ±5.1
1,000
12
21.9 ±0.3
15.3 ±1.3
14.5 ±1.4
14.3 ±1.4
14.2 ±1.6(1)
103/703
93.6 ±6.1
5,000
12
22.6±0.5C
14.6 ±l.la
13.1±2.2a
12.6±2.5b
10.3 ± 2.8 (9)c
75/81
86.2±16.2a
     ap < 0.05, bp < 0.01, °p < 0.001, as calculated by the authors.
     Values are means ± S.D. Values in parentheses are the numbers of litters.
     Source: NEDO(1987).

 1
 2          NEDO (1987) contains an account of a two-generation reproductive study that evaluated
 3   the effects of pre- and postnatal methanol (reagent grade) exposure (20 hours/day) on
 4   reproductive and other organ systems of Sprague-Dawley rats. The F0 generation (30 males and
 5   30 females per exposure group)37 was exposed to 0, 10, 100, and 1,000 ppm (0, 13.1, 131, and
 6   1,310 mg/m3) from 8 weeks old to the end of mating (males) or to the end of lactation period
 7   (females).  The FI generation was exposed to the same concentrations from birth to the end of
 8   mating (males) or to weaning of F2 pups 21 days after delivery (females). Males and females of
 9   the F2 generation were exposed from birth to 21  days old (one animal/sex/litter was exposed to
10   8 weeks of age). NEDO (1987) noted reduced brain, pituitary, and thymus weights, and early
11   testicular descent in the offspring of F0 and FI rats exposed to 1,000 ppm methanol.  The early
12   testicular descent is believed to be an indication  of earlier fetal development as indicated by the
13   fact that it was correlated with increased pup body weight. However, no histopathologic effects
14   of methanol were observed.  As discussed in the report, NEDO (1987) sought to confirm the
15   possible compound-related effect of methanol on the brain by carrying out an additional study in
16   which Sprague-Dawley rats were exposed to 0, 500, 1,000, and 2,000 ppm (0, 655, 1,310, and
17   2,620 mg/m3) methanol from the first day of gestation through the FI generation (see
18   Section 4.4.2).
       A second control group of 30 animals/sex was maintained in a separate room to "confirm that environmental conditions inside
     the chambers were not unacceptable to the animals."
                                               4-37
DRAFT-DO NOT CITE OR QUOTE

-------
 1          Rogers et al. (1993a) evaluated development toxicity in pregnant female CD-I mice
 2    exposed to air or 1,000, 2,000, 5,000, 7,500, 10,000, or 15,000 ppm (0, 1,310, 2,620, 6,552,
 3    9,894, 13,104, and 19,656 mg/m3) methanol vapors (> 99.9% purity) in a chamber for
 4    7 hours/day on GD6-GD15 in a 3-block design experiment.  The numbers of mice exposed at
 5    each dose were 114, 40, 80, 79, 30, 30, and 44, respectively.  During chamber exposures to air or
 6    methanol, the mice had access to water but not food. In order to determine the effects of the
 7    chamber exposure conditions, an additional 88 control mice were not handled and remained in
 8    their cages; 30 control mice were not handled but were food  deprived for 7 hours/day on GD6-
 9    GDIS.  Effects in dams and litters were statistically analyzed using the General Linear Models
10    procedure and multiple t-tesi of least squares means for continuous variables and the Fisher's
11    exact test for dichotomous variables. An analysis of plasma methanol levels in 3 pregnant
12    mice/block/treatment group on GD6, GD10, and GDIS revealed a dose-related increase in
13    plasma methanol concentration that did not seem to reach saturation levels, and methanol plasma
14    levels were not affected by gestation stage or number of previous exposure days. Across all
15    3 days, the mean plasma methanol concentrations in pregnant mice were approximately 97, 537,
16    1,650, 3,178, 4,204, and 7,330 |ig/mL in the 1,000, 2,000, 5,000,  7,500, 10,000, and 15,000 ppm
17    exposure groups, respectively.
18          The dams exposed to air or methanol in chambers gained  significantly less weight than
19    control dams that remained in cages and were  not handled. There were no methanol-related
20    reductions in maternal body weight gain or overt signs  of toxicity. Dams were sacrificed on
21    GDI 7 for a comparison of developmental toxicity in methanol-treated groups versus the chamber
22    air-exposed control group. Fetuses in all exposure groups were weighed, assessed for viability,
23    and examined for external malformations. Fetuses in the control, 1,000, 2,000, 5,000, and
24    15,000 ppm groups were also examined for skeletal and visceral defects. Incidence of
25    developmental effects is listed in Table 4-8. A statistically significant increase in cervical
26    ribs/litter was observed at concentrations of 2,000, 5,000, and 15,000 ppm. At doses of
27    > 5,000 ppm the incidences of cleft palates/litter and exencephaly/litter were increased with
28    statistical significance achieved at all concentrations with the exception of exencephaly which
29    increased but not significantly at 7,500 ppm.38 A significant  reduction in live pups/litter was
30    noted at >7,500 ppm, with a significant increase in fully resorbed litters occurring at
31    > 10,000 ppm. Fetal weight was significantly reduced at > 10,000 ppm. Rogers et al. (1993a)
32    identified a developmental NOAEL and LOAEL of 1,000 ppm and 2,000 ppm, respectively.
33    They also provide BMD maximum likelihood  estimates (benchmark concentration [BMC];
34    referred  to by the authors as MLE) and estimates of the lower 95% confidence limit on the BMC
      38 Due to the serious nature of this response and the relative lack of a response in controls, all incidence of
      exenceaphaly reported in this study at 5,000 ppm or higher are considered biologically significant.
                                               4-3 8        DRAFT-DO NOT CITE OR QUOTE

-------
1    (benchmark concentration, 95% lower bound [BMCL]; referred to as benchmark dose [BMD] by
2    Rogers et al. (1993a) for 5% and 1% added risk, by applying a log-logistic dose-response model
3    to the mean percent/litter data for cleft palate, exencephaly and resorption. The BMCos and
4    BMCLos values for added risk estimated by Rogers et al. (1993a) are listed in Table 4-9. From
5    this analysis, the most sensitive indicator of developmental toxicity was an increase in the
6    proportion of fetuses per litter with cervical rib anomalies.  The most sensitive BMCL and BMC
7    from this effect for 5% added risk were 305 ppm (400 mg/m3) and 824 ppm (1,080 mg/m3),
8    respectively.39

         Table 4-8. Developmental effects in mice after methanol inhalation
Endpoint
No. live pups/litter
No. fully resorbed litters
Fetus weight (g)
Cleft palate/ litter (%)
Exencephaly /litter (%)
Exposure concentration (ppm)
0
9.9
0
1.20
0.21
0
1,000
9.5
0
1.19
0.65
0
2,000
12.0
0
1.15
0.17
0.88
5,000
9.2
0
1.15
8.8b
6.9a
7,500
8.6b
3
1.17
46.6C
6.8
10,000
7.3C
5a
1.04C
52.7C
27.4C
15,000
2.2C
14C
0.70C
48.3C
43.3C
Anomalies
Cervical ribs/litter (%)
Sternebral defects/litter (%)
Xiphoid defects/litter (%)
Vertebral arch defects/litter (%)
Extra lumbar ribs/litter (%)
28
6.4
6.4
0.3
8.7
33.6
7.9
3.8
ND
2.5
49.6b
3.5
4.1
ND
9.6
74.4C
20.2C
10.9
1.5
15.6
ND
ND
ND
ND
ND
ND
ND
ND
ND
ND
60.0a
100C
73.3C
33.3C
40.0C
Ossifications (values are means of litter means)
Sternal
Caudal
Metacarpal
Proximal phalanges
Metatarsals
Proximal phalanges
Distal phalanges
Supraoccipital score+
5.96
5.93
7.96
7.02
9.87
7.18
9.64
1.40
5.99
6.26
7.92
7.04
9.90
7.69
9.59
1.65
5.94
5.71a
7.96
7.04
9.87
6.91
9.57
1.57
5.81
5.42
7.93
6.12
9.82
5.47
8.46b
1.48
ND
ND
ND
ND
ND
ND
ND
ND
ND
ND
ND
ND
ND
ND
ND
ND
5.07C
3.20a
7.60b
3.33C
8.13C
Oc
4.27C
3.20C
     ND = Not determined.  = on a scale of 1-4, where 1 is fully ossified and 4 is unossified.
     Statistical significance: ap < 0.05, bp < 0.01, cp < 0.001, as calculated by the authors.
     Source: Rogers etal. (1993a).
     39 The BMD analysis of the data described in Section 5 was performed similarly using, among others, a similar
     nested logistic model. However, the Rogers et al. (1993a) analysis was performed using added risk and external
     exposure concentrations, whereas the analyses in Section 5 used extra risk and internal dose metrics that were then
     converted to human equivalent exposure concentrations.
                                                 4-39        DRAFT-DO NOT CITE OR QUOTE

-------
          Table 4-9. Benchmark doses at two added risk levels
Endpoint
Cleft Palate (CP)
Exencephaly (EX)
CPandEX
Resorptions (RES)
CP, EX, and RES
Cervical ribs
BMCos (ppm)
4,314
5,169
3,713
5,650
3,667
824
BMCL05(ppm)
3,398
3,760
3,142
4,865
3,078
305
BMCoi (ppm)
2,717
2,122
2,381
3,749
2,484
302
BMCLoi (ppm)
1,798
784
1,816
2,949
1,915
58
      Source: Rogers etal. (1993a).

 1          Rogers and Mole (1997) investigated the critical period of sensitivity to the
 2    developmental toxicity of inhaled methanol in the CD-I mouse by exposing 12-17 pregnant
 3    females to 0 or 10,000 ppm (0 and 13,104 mg/m3), 7 hours/day on 2 consecutive days during
 4    GD6-GD13, or to a single exposure to the same methanol concentration during GD5-GD9.
 5    Another group of mice received a single 7-hour exposure to methanol at 10,000 ppm. The latter
 6    animals were sacrificed at various time intervals up to 28 hours after exposure.  Blood samples
 7    were taken from these animals to measure the concentration of methanol in the serum. Serum
 8    methanol concentrations peaked at ~4 mg/mL 8 hours after the onset of exposure.  Methanol
 9    concentrations in serum had declined to pre-exposure levels after 24 hours. All mice in the main
10    body of the experiment were sacrificed on GDI 7,  and their uteri removed.  The live, dead, and
11    resorbed fetuses were counted, and all  live fetuses were weighed,  examined externally for cleft
12    palate, and then preserved.  Skeletal abnormalities were determined after the carcasses had been
13    cleaned and  eviscerated. Cleft palate, exencephaly, and skeletal defects were observed in the
14    fetuses of exposed dams. For example, cleft palate was observed following 2-day exposures to
15    methanol on GD6-GD7 through GDI 1-GD12.  These effects also were apparent in mice
16    receiving a single exposure to methanol on GD5-GD9.  This effect peaked when the dams were
17    exposed on GD7.  Exencephaly showed a similar pattern of development in response to methanol
18    exposure. However, the data indicated  that cleft palate and exencephaly might be competing
19    malformations, since only one fetus displayed both features. Skeletal malformations included
20    exoccipital anomalies, atlas and axis defects, the appearance of an extra rudimentary rib on
21    cervical vertebra No.7, and supernumerary lumbar ribs.  In each case, the maximum time point
22    for the induction of these defects appeared to be when the dams were exposed to methanol on or
23    near GD7. When dams were exposed to methanol on GD5, there was also an increased
24    incidence of fetuses with 25 presacral vertebrae (26 is normal). However, an increased incidence
25    of fetuses with 27 presacral vertebrae was evident when dams were exposed on GD7. These
26    results indicate that gastrulation and early organogenesis is a period of increased embryonic
27    sensitivity to methanol.
                                              4-40
DRAFT-DO NOT CITE OR QUOTE

-------
 1          Burbacher et al. (1999a, 1999b) carried out toxicokinetic and reproductive/developmental
 2    studies of methanol in M. fascicularis monkeys that were published by the Health Effects
 3    Institute (HEI) in a two-part monograph.  Some of the data were subsequently published in the
 4    open scientific literature (Burbacher et al., 2004a, 2004b). The experimental protocol featured
 5    exposure to 2 cohorts of 12 monkeys/group to low exposure levels (relative to the previously
 6    discussed rodent studies) of 0, 200, 600, or 1,800 ppm (0, 262, 786, and 2,359 mg/m3) methanol
 7    vapors (99.9% purity), 2.5 hours/day, 7 days/week, during a premating period and mating period
 8    (-180 days combined) and throughout the entire gestation period (-168 days). The monkeys
 9    were 5.5-13 years old and were a mixture of feral-born and colony-bred animals. The study
10    included an evaluation of maternal reproductive performance and tests to assess infant postnatal
11    growth and newborn health, reflexes, behavior, and development of visual, sensorimotor,
12    cognitive, and social behavioral function (see Section 4.4.2 for a review of the developmental
13    neurotoxicity findings from this study). Blood methanol levels, clearance, and the appearance of
14    formate were also examined and are discussed in Section 3.2.
15          With regard to reproductive parameters, there was a statistically significant decrease
16    (p = 0.03) in length of pregnancy in all treatment groups, as shown in Table 4-10. Maternal
17    menstrual cycles, conception rate, and live birth index were all unaffected by exposure. There
18    were also no signs of an effect on maternal weight gain or clinical toxicity among the dams. The
19    decrease in pregnancy length was largely due to complications of pregnancy requiring Cesarean
20    section (C-section) deliveries in the methanol exposure groups.  The C-section deliveries were
21    performed in response to signs of difficulty in the pregnancy and thus may serve as supporting
22    evidence of reproductive dysfunction in the methanol-exposed females.
23          While pregnancy  duration was virtually the same in all exposure groups, there were some
24    indications of increased pregnancy duress only in methanol-exposed monkeys. C-sections were
25    done in 2 monkeys from the 200 ppm group and 2 from the 600 ppm group due to vaginal
26    bleeding, presumed, but not verified, to be from placental detachment.40 A monkey in the
27    1,800 ppm group also received a C-section after experiencing nonproductive labor for 3 nights.
28    In addition, signs of prematurity were observed in 1 infant from the 1,800 ppm group that was
29    born after a 150-day gestation period.  The authors speculated that the shortened gestation length
30    could be due to a direct effect of methanol on the fetal hypothalamus-pituitary-adrenal (HPA)
31    axis or an indirect effect of methanol on the maternal uterine environment. Other fetal
32    parameters such as crown-rump length and head circumference were unchanged among the
33    groups.  Infant growth and tooth eruption were unaffected by prenatal methanol exposure.
      40 Burbacher et al. (2004a, 2004b) note, however, that in studies of pregnancy complication in alcohol- exposed
      human subjects, an increased incidence of uterine bleeding and abrutio placenta has been reported.
                                                4-41         DRAFT-DO NOT CITE OR QUOTE

-------
          Table 4-10. Reproductive parameters in monkeys exposed via inhalation to methanol
          during prebreeding, breeding, and pregnancy
Exposure (ppm)
0
200
600
1,800
Conception rate
9/11
9/12
9/11
10/12
Weight gain (kg)
1.67 ±0.07
1.27 ±0.14
1.78 ±0.25
1.54 ±0.20
Pregnancy duration (days)3
168 ±2
160 ± 2b
162 ± 2b
162 ± 2b
Live born delivery rate
8/9
9/9
8/9
9/10
      Values are means ± SE.;
      aLive-born offspring only; bp < 0.05, as calculated by the authors.
      Source: Burbacher et al. (2004a).
 1          In later life, 2 females out of the total of 9 offspring in the 1,800 ppm group experienced
 2    a wasting syndrome at 12 and 17 months of age. Food intake was normal and no cause of the
 3    syndrome could be determined in tests for viruses, hematology, blood chemistry, and liver,
 4    kidney, thyroid, and pancreas function.  Necropsies revealed gastroenteritis and severe
 5    malnourishment. No infectious agent or other pathogenic factor could be identified. Thus, it
 6    appears that a highly significant toxicological effect on postnatal growth can be attributed to
 7    prenatal methanol  exposure at 1,800 ppm (2,300 mg/m3).
 8          In summary, the Burbacher et al. (2004a, 2004b,  1999b) studies suggest that methanol
 9    exposure can cause reproductive effects, manifested as a shortened mean gestational period due
10    to pregnancy complications that precipitated delivery via a C-section, and developmental
11    neurobehavioral effects which may be related to the shortened gestational period (see
12    Section 4.4.2). The low exposure of 200 ppm may signify a LOAEL for reproductive effects.
13    However, the decrease in gestational length was marginally significant and largely the result of
14    human intervention (C-section) for reasons (presumably pregnancy complications) that were not
15    objectively confirmed with clinical procedures (e.g., placental ultrasound). Also, this effect did
16    not appear to be dose related, the greatest gestational period decrease having occurred at the
17    lowest (200 ppm) exposure level. Thus, a clear NOAEL or LOAEL cannot be determined from
18    this study.
19          In a study of the testicular effects of methanol, Cameron et al.  (1984) exposed 5 male
20    Sprague-Dawley rats/group to methanol vapor, 8 hours/day, 5 days/week for 1, 2, 4, and 6 weeks
21    at 0, 200, 2,000, or 10,000 ppm (0, 262, 2,620, and 13,104 mg/m3). The authors examined the
22    possible effects of methanol  on testicular function by measuring blood levels of testosterone,
23    luteinizing hormone (LH), and follicular stimulating hormone (FSH) using radioimmunoassay.
24    When the authors tabulated their results as a percentage  of the control value for each duration
25    series, the most significant changes were in blood testosterone levels of animals exposed to
26    200 ppm methanol, the lowest concentration evaluated.  At this exposure level, animals exposed
27    for 6 weeks had testosterone levels that were 32% of those seen in controls. However, higher
                                               4-42
DRAFT-DO NOT CITE OR QUOTE

-------
 1   concentrations of methanol were associated with testosterone levels that were closer to those of
 2   controls. However, the lack of a clear dose-response is not necessarily an indication that the
 3   effect is not related to methanol. The higher concentrations of methanol could be causing other
 4   effects (e.g., liver toxicity) which can influence the results. Male rats exposed to 10,000 ppm
 5   methanol for 6 weeks displayed blood levels of LH that were about 3 times higher (mean ± S.D.)
 6   than those exposed to air (311 ± 107% versus 100 ± 23%). In discussing their results, the authors
 7   placed the greater emphasis on the fact that an exposure level equal to the ACGIH TLV
 8   (200 ppm) had caused a significant depression in testosterone formation in male rats.
 9          A follow-up study report by the same research group (Cameron et al.,  1985) described the
10   exposure of 5 male Sprague-Dawley rats/group, 6 hours/day for either 1  day or 1 week, to
11   methanol, ethanol, n-propanol, or n-butanol at their respective TLVs. Groups of animals were
12   sacrificed immediately after exposure or after an 18-hour recovery period, and the levels of
13   testosterone, LH, and corticosterone  measured in serum.  As shown in Table 4-11, the data were
14   consistent with the ability of these aliphatic alcohols to cause a transient reduction in the
15   formation of testosterone.  Except in the case of n-butanol, rapid recovery from these deficits can
16   be inferred from the 18-hour postexposure data.

           Table 4-11. Mean serum levels of testosterone, luteinizing hormone, and
           corticosterone (± S.D.) in male  Sprague-Dawley rats after inhalation of methanol,
           ethanol, n-propanol or n-butanol at threshold limit values
Testosterone (as a percentage of control)
Condition
Control
Methanol
Ethanol
n-Propanol
n-Butanol
TLV
(ppm)

200
1,000
200
50
Single-day exposure
End of exposure
100 ±17
41±16a
64 ± 12a
58±15a
37±8a
18 hr
postexposure
100 ± 20
98 ±18
86 ±16
81 ±13
52 ± 22a
One-week exposure
End of exposure
100 ± 26
81 ±22
88 ±14
106 ± 28
73 ±34
18 hr
postexposure
100 ± 17
82 ±27
101 ±13
89 ±17
83 ±18
Luteinizing hormone
Control
Methanol
Ethanol
n-Propanol
n-Butanol

200
1,000
200
50
100 ±30
86 ±32
110 ±22
117±59
124 ±37
100 ±35
110 ±40
119±54
119±83
115±28
100 ± 28
78 ±13
62 ±26
68 ±22
78 ±26
100 ± 36
70 ±14
81 ±17
96 ±28
98 ±23
                                               4-43
DRAFT-DO NOT CITE OR QUOTE

-------
Testosterone (as a percentage of control)
Condition
TLV
(ppm)
Single-day exposure
End of exposure
18 hr
postexposure
One-week exposure
End of exposure
18 hr
postexposure
Corticosterone
Control
Methanol
Ethanol
n-Propanol
n-Butanol

200
1,000
200
50
100 ± 20
115±18
111±32
112±21
143 ±Ha
ND
ND
ND
ND
ND
100 ±21
74 ±26
60 ±25
79 ±14
85 ±26
ND
ND
ND
ND
ND
      ND = Nodata.;
      ap < 0.05, as calculated by the authors.
      Source:  Cameron etal. (1985).
 2           In a series of studies that are relevant to the reproductive toxicity of methanol in males,
 3    Lee et al. (1991) exposed 8-week-old male Sprague-Dawley rats (9-10/group) to 0 or 200 ppm
 4    (0 and 262 mg/m3) methanol, 8 hours/day, 5 days/week, for 1, 2, 4, or 6 weeks to measure the
 5    possible treatment effects on testosterone production.  Study results were evaluated by one factor
 6    ANOVA followed by Student's ^-test. In the treated rats, there was no effect on serum
 7    testosterone levels, gross structure of reproductive organs, or weight of testes and seminal
 8    vesicles.  Lee et al. (1991) also studied the in vitro effect of methanol on testosterone production
 9    from isolated testes, but saw no effect on testosterone formation either with or without the
10    addition of human chorionic gonadotropin hormone.
11           In a third experiment from the same  report, Lee et al. (1991) examined testicular
12    histopathology to determine if methanol exposure produced lesions indicative of changing
13    testosterone levels; the effects of age and folate status were also assessed. This is relevant to the
14    potential toxicity of methanol because folate is the coenzyme of tetrahydrofolate synthetase, an
15    enzyme that is rate limiting in the removal of formate.  Folate deficiency would be expected to
16    cause potentially toxic levels of methanol, formaldehyde, and formate to be retained. The same
17    authors examined the relevance of folate levels, and by implication, the overall status of formate
18    formation and clearance in mediating the testicular functions of Long-Evans rats.  Groups of
19    4-week-old male Long-Evans rats were given diets containing either adequate or reduced folate
20    levels plus 1% succinylsulfathiazole, an antibiotic that, among other activities,41 would tend to
21    reduce  the folate body burden.  At least 9 rats/dietary group/dose were exposed to 0,  50, 200, or
22    800 ppm (0, 66, 262, and 1,048 mg/m3) methanol vapors starting at 7 months of age  while 8-
23    12 rats/dietary group/dose were exposed to 0 or 800 ppm methanol vapors at 15  months of age.
      41 Succinylsulfathiazone antibiotic may have a direct impact on the effects being measured, the extent of which was
      not addressed by the authors of this study.
                                                4-44
DRAFT-DO NOT CITE OR QUOTE

-------
 1    The methanol exposures were conducted continuously for 20 hours/day for 13 weeks. Without
 2    providing details, the study authors reported that visual toxicity and acidosis developed in rats
 3    fed the low folate diet and exposed to methanol. No methanol-related testicular lesions or
 4    changes in testes or body weight occurred in rats that were fed either the folate sufficient or
 5    deficient diets and were 10 months old at the end of treatment. Likewise, no methanol-lesions
 6    were observed in 18-month-old rats that were fed diets with adequate folate. However, the
 7    incidence but not severity of age-related testicular lesions was increased in the  18-month-old rats
 8    fed folate-deficient diets.  Subcapsular vacuoles in germinal epithelium were noted in 3/12
 9    control rats and  8/13 rats in the 800 ppm group. One rat in the 800 ppm group had atrophied
10    seminiferous tubules and another had Ley dig cell hyperplasia. These effects, as well as the
11    transient decrease in testosterone levels observed by Cameron et al. (1985, 1984), could be the
12    result of chemically-related strain on the rat system as it attempts to maintain hormone
13    homeostasis.
14          Dorman  et al. (1995) conducted a series of in vitro and in vivo studies of developmental
15    toxicity  in ICR BR (CD-I) mice associated with methanol and formate exposure. The studies
16    used HPLC grade methanol and appropriate controls. PK and developmental toxicity parameters
17    were measured in mice exposed to a 6-hour methanol inhalation (10,000 or 15,000 ppm),
18    methanol gavage (1.5 g/kg) or sodium formate  (750 mg/kg by gavage) on GD8. In the in vivo
19    inhalation study, 12-14 dams/group were exposed to 10,000 ppm methanol for 6 hours on
20    GD8,42 with and without the administration of fomepizole (4-methylpyrazole) to inhibit the
21    metabolism of methanol by ADH1. Dams  were sacrificed on GD10, and folate levels in
22    maternal RBC and conceptus (decidual swelling) were measured, as well as fetal neural  tube
23    patency  (an early indicator of methanol-induced dysmorphogenic response). The effects
24    observed included a transient decrease in maternal RBC and conceptus folate levels within
25    2 hours following exposure and a significant (p < 0.05) increase in the incidence of fetuses with
26    open neural tubes (9.65% in treated versus 0 in control).  These responses were not observed
27    following sodium formate administration, despite peak formate levels in plasma and decidual
28    swellings being  similar to those observed following the 6-hour methanol inhalation of
29    15,000 ppm.  This suggests that these methanol-induced effects are not related to the
30    accumulation of formate.  As this study provides information relevant to the identification of the
31    proximate teratogen associated with developmental toxicity in rodents, it is discussed more
32    extensively in Section 4.6.1.
      42 Dorman et al. (1985) state that GD8 was chosen because it encompasses the period of murine neurulation and the
      time of greatest vulnerability to methanol-induced neural tube defects.
                                                4-45        DRAFT-DO NOT CITE OR QUOTE

-------
      4.3.3.  Other Reproductive and Developmental Toxicity Studies
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
            Additional information relevant to the possible effects of methanol on reproductive and
      developmental parameters has been provided by experimental studies that have exposed
      experimental animals to methanol during pregnancy via i.p. injections (Rogers et al., 2004).
      Relevant to the developmental impacts of the chemical, a number of studies also have examined
      the effects of methanol when included in whole-embryo culture (Hansen et al., 2005; Harris
      et al., 2003; Andrews etal., 1998, 1995, 1993).
            Pregnant female C57BL/6J mice received 2 i.p. injections  of methanol on GD7 (Rogers
      et al., 2004).  The injections were given 4 hours apart to provide a total dosage of 0, 3.4, and 4.9
      g/kg. Animals were sacrificed on GDI7 and the litters were examined for live,  dead, and
      resorbed fetuses.  Rogers et al. (2004) monitored fetal weight and examined the fetuses for
      external abnormalities and skeletal malformations.  Methanol-related deficits in maternal and
      litter parameters observed by Rogers et al. (2004) are summarized in Table 4-12.

          Table 4-12. Maternal and litter parameters when pregnant female C57BL/6J mice
          were injected i.p. with methanol
Parameter
No. pregnant at term
WtgainGD7-GD8(g)
WtgainGD7-GD10(g)
Live fetuses/litter
Resorbed fetuses/litter
Dead fetuses/litter
Fetal weight (g)
Methanol dose (g/kg)
0
43
0.33 ±0.10
1.63 ±0.18
7.5 ±0.30
0.4 ±0.1
0.1±0.1
0.83 ±0.02
3.4
13
0.37 ±0.15
2.20 ± 0.20
6.3±0.5a
1.3±0.4a
0
0.82 ±0.03
4.9
24
-0.24±0.14a
1.50 ±0.20
3.7±0.4a
4.4±0.4a
0.1±0.1
0.70 ± 0.02a
13
14
15
16
17
18
19
20
21
     Values are means ± SEM.
     ap < 0.05, as calculated by the authors.
     Source: Rogers et al. (2004).

            Rogers et al. (2004) used a number of sophisticated imaging techniques, such as confocal
     laser scanning and fluorescence microscopy, to examine the morphology of fetuses excised at
     GD7, GD8, and GD9.  They identified a number of external craniofacial abnormalities, the
     incidence of which was, in all cases, significantly increased in the high-dose group compared to
     controls. For some responses, such as microanophthalmia and malformed maxilla, the incidence
     was also significantly increased in animals receiving the lower dose. Fifteen compound-related
     skeletal malformations were tabulated in the report. In most cases, a dose-response effect was
     evident, resulting in statistically significant incidences in affected fetuses and litters, when
     compared to controls. Apparent effects of methanol on the embryonic forebrain included a
                                               4-46
                                                           DRAFT-DO NOT CITE OR QUOTE

-------
 1    narrowing of the anterior neural plate, missing optical vesicles, and holoprosencephaly (failure of
 2    the embryonic forebrain to divide). The authors noted that there was no sign of incipient cleft
 3    palate or exencephaly, as had been observed in CD-I mice exposed to methanol via the oral and
 4    inhalation routes (Rogers et al.,  1993a).
 5          In order to collect additional information on cell proliferation and histological changes in
 6    methanol-treated fetuses, Degitz et al. (2004a) used an identical experimental protocol to that of
 7    Rogers et al. (2004) by administering 0, 3.4, or 4.9 g methanol/kg in distilled water i.p. (split
 8    doses, 4 hours apart) to C57BL/6J mice on GD7. Embryos were collected at various times on
 9    GD8 and GD10. Embryos from dams exposed to 4.9 g/kg and examined on GD8 exhibited
10    reductions in the anterior mesenchyme, the mesenchyme subjacent to the mesencephalon and the
11    base of the prosencephalon (embryonic forebrain), and in the forebrain epithelium. The optic
12    pits were often lacking;  where present their epithelium was thin and there were fewer neural
13    crest cells in the mid- and hindbrain regions.
14          At GD9, there was  extensive cell death in areas populated by the neural crest, including
15    the forming cranial ganglia. Dose-related abnormalities in the development of the cranial nerves
16    and ganglia were seen on GD7.  In accordance with an arbitrary dichotomous scale devised by
17    the authors, scores for ganglia V, VIII, and IX were significantly (not otherwise specified)
18    reduced at all dose levels, and ganglia VII and X were reduced only at the highest dose.  At the
19    highest dose (4.9 g/kg), the brain and face were poorly developed and the brachial  arches were
20    reduced in size or virtually absent. Flow cytometry of the head regions of the embryos from the
21    highest dose at GD8 did not show an effect on the proportion of cells in S-phase.
22          Cell growth  and development were compared in C57BL/6J and CD-I mouse embryos
23    cultured in methanol (Degitz, et al., 2004b).  GD8 embryos, with 5-7 somites, were cultured in
24    0, 1, 2, 3, 4, or 6 mg methanol/mL for 24 hours and evaluated for morphological development.
25    Cell death was increased in both strains in a  developmental stage- and region-specific manner at
26    4 and 6 mg/mL after 8 hours of exposure.  The proportions of cranial region cells in S-phase
27    were significantly (p < 0.05) decreased at 6 mg/mL following 8- and 18-hour exposures to
28    methanol. After 24 hours of exposure, C57BL/6J embryos had significantly (p < 0.05) decreased
29    total protein at 4 and 6 mg/kg.  Significant (p < 0.05) developmental effects were seen at 3, 4,
30    and 6 mg/kg, with eye dysmorphology being the most sensitive endpoint. CD-I embryos had
31    significantly decreased total protein at 3, 4, and 6 mg/kg, but developmental effects were seen
32    only at 6 mg/kg. It was concluded that the C57BL/6J embryos were more severely affected by
33    methanol in culture  than the CD-I  embryos.
34          Andrews et al. (1993) carried out a comparative study of the developmental toxicity of
35    methanol in whole Sprague-Dawley rat or CD-I mouse embryos.  Nine-day rat embryos were
36    explanted and cultured in rat serum containing 0, 2, 4, 8, 12, or 16 mg/mL methanol for 24 hours

                                               4-47        DRAFT-DO NOT CITE OR QUOTE

-------
 1   then transferred to rat serum alone for a further 24 hours. Eight-day mouse embryos were
 2   cultured in 0, 2, 4, 6, or 8 mg/mL methanol in culture medium for 24 hours.  At the end of the
 3   culture period, embryos were examined for growth, development and dysmorphogenesis.  For
 4   the rats, doses of 8 mg/mL and above resulted in a concentration-related decrease in somite
 5   number, head length, and developmental score. Some lethality was seen in embryos incubated at
 6   12 mg/mL methanol.  For the mouse embryos, incubation concentrations of 4 mg/mL methanol
 7   and above resulted in a significant decrease in developmental score and crown-rump length. The
 8   high concentration (8 mg/mL) was associated with embryo lethality. These data  suggest that
 9   mouse embryos are more sensitive than rat embryos to the developmental effects of methanol.
10   Using a similar experimental system to examine the developmental toxicity of formate and
11   formic acid in comparison to methanol, Andrews et al. (1995) showed that the formates are
12   embryotoxic at doses that are four times lower than equimolar doses of methanol. Andrews et al.
13   (1998) showed that exposure to combinations of methanol and formate was less embryotoxic
14   than would be expected based on simple toxicity additivity, suggesting that the embryotoxicity
15   observed following low-level exposure to methanol is mechanistically different from that
16   observed following exposure to formate.
17          A study by Hansen et al. (2005) determined the comparative toxicity of methanol and its
18   metabolites, formaldehyde and sodium formate, in GD8 mouse (CD-I) and GD10 rat (Sprague-
19   Dawley) conceptuses.  Incubation of whole embryos was for 24 hours in chemical-containing
20   media (mouse: 4-12 mg/mL methanol, 1-6 |ig/mL formaldehyde, 0.5-4 mg/mL sodium formate;
21   rat: 8-20 mg/mL, 1-8  jig/mL, 0.5-8 mg/mL).  Subsequently, the visceral yolk sac (VYS) was
22   removed and frozen for future protein and DNA determination.  The embryos were examined
23   morphologically to determine growth and developmental parameters such as viability, flexure
24   and rotation, crown-rump length, and neuropore closure. In other experiments, the chemicals
25   were injected directly into the amniotic space.  For each response, Table 4-13 provides a
26   comparison of the concentrations or amounts of methanol, formaldehyde,  and formate that
27   resulted in statistically significant changes in developmental abnormalities compared to controls.
28          For a first approximation, these concentrations or amounts may be taken as threshold-
29   dose ranges for the specific responses under the operative experimental conditions. The data
30   show consistently lower threshold values for the effects of formaldehyde compared to those of
31   formate and methanol. The mouse embryos were more sensitive towards methanol toxicity than
32   rat embryos, consistent with in vivo findings, whereas the difference in  sensitivity disappeared
33   when formaldehyde was administered. Hansen et al.  (2005) hypothesized that, while the MOA
34   for the initiation of the organogenic defects is unknown, the relatively low threshold levels of
35   formaldehyde for most measured effects suggest formaldehyde involvement in the embryotoxic
36   effects of methanol.  By contrast, formate, the putative toxicant for the acute effects of methanol

                                              4-48        DRAFT-DO NOT CITE OR QUOTE

-------
 1   poisoning (acidosis, neurological deficits), did not appear to reproduce the methanol-induced
 2   teratogenicity in these whole embryo culture experiments.

          Table 4-13. Reported thresholds concentrations (and author-estimated ranges) for the
          onset of embryotoxic effects when rat and mouse conceptuses were incubated in vitro
          with methanol, formaldehyde, and formate
Parameter
Viability (%)
Normal rotation (%)
CRa length
Neural tube closure (%)
Reduced embryo protein
Reduced VYSb protein
Reduced embryo DNA
Reduced VYS DNA
Mouse
Methanol
Formaldehyde
Formate
Rat
Methanol
Formaldehyde
Formate
In vitro incubation (mg/mL)
8.0
4.0
No change
8.0
8.0
10.0
8.0
4.0
0.004
0.003
No change
0.001
0.003
0.004
0.003
0.001
NS
0.5
No change
2.0
4.0
4.0
No change
0.5
16.0
8.0
No change
12.0
8.0
12.0
12.0
12.0
0.006
0.003
No change
No change
0.004
0.004
0.003
0.003
2.0
4.0
No change
No change
2.0
NR
NR
NR
Microinjection (author-estimated dose ranges in jig)
Viability (%)
Normal rotation (%)
CRa length
Neural tube closure (%)
Reduced embryo protein
Reduced VYSb protein
Reduced embryo DNA
Reduced VYSb DNA
46-89
1-45
No change
1-45
1-45
135-178
46-89
1-45
0.003-0.5
0.003-0.5
No change
0.003-0.5
0.501-1.0
1.01-1.5
0.501-1.0
0.003-0.5
1.01-1.5
0.03-0.5
No change
1.01-1.5
No change
No change
No change
0.03-0.5
46-89
46-89
No change
No change
No change
No change
No change
No change
1.01-1.5
1.01-1.5
No change
No change
1.51-2.0
No change
No change
No change
1.51-4.0
0.51-1.0
No change
1.01-1.5
0.51-1.0
1.01-1.5
0.51-1.0
0.51-1.0
 4
 5
 6
 7
 8
 9
10
11
12
aCR = crown-rump length,
bVYS = visceral yolk sac.
NR = not reported
Source: Hansen et al. (2005); Harris et al. (2004) (adapted).

      Harris et al. (2003) provided biochemical evidence consistent with the concept that
formaldehyde might be the ultimate embryotoxicant of methanol by measuring the activities of
enzymes that are involved in methanol metabolism in mouse (CD-I) and rat (Sprague-Dawley)
whole embryos  at different stages of development.  Specific activities of the enzymes ADH1,
ADH3, and CAT, were determined in rat and mouse conceptuses during the organogenesis period
of 8-25 somites. Activities were measured in heads, hearts, trunks, and VYS from early- and
late-stage mouse and rat embryos. While CAT activities were similar between rat and mouse
embryos, mouse ADH1 activities in the VYS were significantly lower throughout organogenesis
when compared to the rat VYS or embryos of either species. ADH1 activities of heads, hearts,
and trunks from mouse embryos were significantly lower than those from rats at the 7-12 somite
                                              4-49
                                                     DRAFT-DO NOT CITE OR QUOTE

-------
 1    stage.  However, these interspecies differences were not evident in embryos of 20-22 somites.
 2    ADH3 activities were lower in mouse versus rat VYS, irrespective of the stage of development.
 3    However, while ADH3 activities in mouse embryos were markedly lower than those of rats in
 4    the early stages of development, the levels of activity were similar to at the 14-16 somite stage
 5    and beyond. A lower capacity to transform formaldehyde to formate might explain the increased
 6    susceptibility of mouse versus rat embryos to the toxic effects of methanol. The hypothesis that
 7    formaldehyde is the ultimate embryotoxicant of methanol is  supported by the demonstration of
 8    diminished ADH3 activity in mouse versus rat embryos and  by the demonstration by Hansen
 9    et al. (2005) that formaldehyde has a far greater embryotoxicity than either formate or methanol
10    itself.
11          That formate can induce similar developmental lesions in whole rat and mouse
12    conceptuses was demonstrated by Andrews et al. (1995), who evaluated the developmental
13    effects of sodium formate and formic acid in rodent whole embryo cultures in vitro. Day 9 rat
14    (Sprague-Dawley) embryos were cultured for 24 or 48 hours and day 8 mouse (CD-I) cultures
15    were incubated for 24 hours. As tabulated by the authors, embryos of either  species showed
16    trends towards increasing lethality and incidence of abnormalities with exposure concentration.
17    Among the anomalies observed were open anterior and posterior neuropores, plus rotational
18    defects, tail anomalies, enlarged pericardium, and delayed heart development.

      4.4. NEUROTOXICITY
19          A substantial body of information exists on  the toxicological consequences to humans
20    who consume or are exposed to large amounts of methanol.  As discussed in Section 4.1,
21    neurological consequences of acute methanol intoxication in humans include Parkinson-like
22    responses, visual impairment, confusion, headache, and numerous subjective symptoms. The
23    occurrence of these symptoms has been shown to be associated with necrosis of the putamen
24    when neuroimaging techniques have been applied (Salzman, 2006).  Such profound changes
25    have been linked to tissue acidosis that arises when methanol is metabolized to formaldehyde
26    and formic acid through the actions of ADH1 and ADH3.  However, the well-documented impact
27    of the substantial amounts of formate that are formed when humans and animals are exposed to
28    large amounts of methanol may obscure the potentially harmful effects that may arise when
29    humans and animals exposed to smaller amounts. Human acute exposure studies (Chuwers et al.,
30    1995; Cooketal., 1991) (See Section 4.1.3) at TLV levels of 200 ppm  would indicate that some
31    measures of neurological function (e.g., sensory evoked potentials, memory testing and
32    psychomotor testing) were impaired in the absence of measureable formate production.
                                              4-50        DRAFT-DO NOT CITE OR QUOTE

-------
      4.4.1. Oral Studies
 1          Two rodent studies investigated the neurological effects of developmental methanol
 2    exposure via the oral route (Aziz et al., 2002; Infurna and Weiss, 1986). One of these studies also
 3    investigated the influence of FAD diets on the effects of methanol exposures (Aziz et al., 2002).
 4    In the first, Infurna and Weiss (1986) exposed 10 pregnant female Long-Evans rats/dose to 2%
 5    methanol (purity not specified) in drinking water on either GDIS-GDI? or GD17-GD19.  Daily
 6    methanol intake was calculated at 2,500 mg/kg-day by the study authors.  Dams were allowed to
 7    litter and nurse their pups. Data were analyzed by ANOVA with the litter as the statistical  unit.
 8    Results of the study were equivalent for both exposure periods. Treatment had no effect on
 9    gestational length or maternal bodyweight. Methanol had no effect on maternal behavior as
10    assessed by the time it took dams to retrieve pups after they were returned to the cage following
11    weighing.  Litter size, pup birth weight, pup postnatal weight gain, postnatal mortality, and day
12    of eye opening did not differ from controls in the methanol treated groups. Two neurobehavioral
13    tests were conducted in offspring.  Suckling ability was tested in 3-5 pups/treatment group on
14    PND1.  An increase in the mean latency for nipple attachment was observed in pups from the
15    methanol treatment group, but the percentage of pups that successfully attached to nipples did
16    not differ significantly between treatment groups.  Homing behavior, the ability to detect home
17    nesting material within a cage  containing one square of shavings from the pup's home cage and
18    four squares of clean shavings, was evaluated in 8  pups/group on PND10.  Pups from both of the
19    methanol exposure groups took about twice as long to locate the home material and took less
20    direct paths than the control pups.  Group-specific values differed significantly from controls.
21    This study suggests that developmental toxicity can occur at this drinking water dose without
22    readily apparent signs of maternal toxicity.
23          Aziz et al. (2002) investigated the role of developmental deficiency in  folic acid and
24    methanol-induced developmental neurotoxicity in PND45 rat pups.  Wistar albino female rats
25    (80/group) were fed FAD43 and FAS diets separately. Following 14-16 weeks on the diets, liver
26    folate levels were estimated and females exhibiting a significantly low folic acid level were
27    mated.  Throughout their lactation period, dams of both the FAD and the FAS  group were given
28    0,1, 2, or 4% v/v methanol via drinking water, equivalent to approximately 480, 960 and
29    1,920 mg/kg-day.44 Pups were exposed to methanol via lactation from PND1-PND21. Litter size
30    was culled to  8 with equal male/female ratios maintained as much as possible. Liver folate
31    levels were determined at PND21 and neurobehavioral parameters (motor performance using the
      43
        Along with the FAD diet, 1% succinylsulphathiazole was also given to inhibit folic acid biosynthesis from
      intestinal bacteria.
      44 Assuming that Wistar rat drinking water consumption is 60 mL/kg-day (Rogers et al., 2002), 1% methanol in
      drinking water would be equivalent to 1% x 0.8 g/mL x 60 mL/kg-day = 0.48 g/kg-day = 480 mg/kg-day.
                                                4-51        DRAFT-DO NOT CITE OR QUOTE

-------
 1    spontaneous locomotor activity test and cognitive performance using the conditioned avoidance
 2    response [CAR] test), and neurochemical parameters (dopaminergic and cholinergic receptor
 3    binding and dopamine levels) were measured at PND45. The expression of growth-associated
 4    protein (GAP 43), a neuro-specific protein in the hippocampus that is primarily localized in
 5    growth cone membranes and is expressed during developmental regenerative neurite outgrowth,
 6    was examined using immunohistochemistry  and western blot analysis.
 7          A loss in body weight gain was observed at PND7, PND14, and PND21 in animals
 8    exposed to 2% (11,  15 and 19% weight gain  reduction) and 4% (17, 24 and 29% weight gain
 9    reduction) methanol in the FAD group and only at 4% (9, 14 and 17% weight gain reduction)
10    methanol in the FAS group. No significant differences in food and water intake were observed
11    among the different treatment groups.  Liver folate levels in the FAD group were decreased by
12    63% in rats prior to mating and 67% in pups on PND21.
13          Based on reports of Parkinson-like symptoms in survivors of severe methanol poisoning
14    (see Section 4.1), Aziz et al. (2002) hypothesized that methanol may cause a depletion in
15    dopamine levels and degeneration of the dopaminergic nigrostriatal pathway.45  Consistent with
16    this hypothesis, they found dopamine levels were significantly decreased (32% and 51%) in the
17    striatum of rats in the FAD group treated with 2% and 4% methanol, respectively. In the FAS
18    group, a significant decrease (32%) was observed in the 4% methanol-exposed  group.
19          Methanol treatment at 2% and 4% was associated with significant increases in activity, in
20    the form of distance traveled in a spontaneous locomotor activity test,  in the FAS group (13%
21    and 39%, respectively) and more notably, in  the FAD group (33% and 66%, respectively) when
22    compared to their respective controls. Aziz et al. (2002) suggest that these alterations in
23    locomotor activity may be caused by a significant alteration in dopamine receptors and
24    disruption in neurotransmitter availability. Dopamine receptor (D2) binding in the hippocampus
25    of the FAD group was significantly increased (34%) at 1% methanol, but was significantly
26    decreased at 2% and 4% methanol exposure  by 20% and 42%, respectively.  In  the FAS group,
27    D2 binding was significantly increased by 22% and 54% in the 2% and 4% methanol-exposed
28    groups.
29          At PND45, the CAR in FAD rats exposed to 2% and 4% methanol was significantly
30    decreased by 48% and 52%, respectively, relative to nonexposed controls. In the FAS group, the
31    CAR was only significantly decreased in the 4% methanol-exposed animals and only by 22% as
32    compared to their respective controls. Aziz et al. (2002) suggest that the impairment in CAR of
33    the methanol-exposed FAD pups may be due to alterations in the number of cholinergic
34    (muscarinic) receptor proteins in the hippocampal region of the brain.  Muscarinic receptor
      45 The nigrostriatal pathway is one of four major dopamine pathways in the brain that are particularly involved in the
      production of movement. Loss of dopamine neurons in the substantia nigra is one of the pathological features of
      Parkinson's disease (Kim et al., 2003),
                                               4-52        DRAFT-DO NOT CITE OR QUOTE

-------
 1   binding was significantly increased in the 2% (20%) and 4% (42%) methanol-exposed group in
 2   FAD animals, while FAS group animals had a significant increase in cholinergic binding only in
 3   the 4% methanol exposed group (21%). High concentrations of methanol may saturate the
 4   body's ability to remove toxic metabolites, including formaldehyde and formate, and this may be
 5   exacerbated in FAD pups having a low store of folate.
 6          Immunohistochemistry showed an increase in the expression of GAP-43 protein in the
 7   dentate granular and pyramidal cells of the hippocampus in 2% and 4% methanol-exposed
 8   animals in the FAD group. The FAS group showed increased expression only in the 4%
 9   methanol-exposed group. The Western blot analysis also confirmed a higher expression of
10   GAP-43 in the 2% and 4% methanol-exposed FAD group rats.  Aziz et al. (2002) suggested that
11   up-regulation of GAP-43 in the hippocampal region may be associated with axonal growth or
12   protection of the nervous system from methanol toxicity.
13          The Aziz et al. (2002) study provides  evidence that hepatic tetrahydrofolate is an
14   important contributing factor in methanol-induced developmental neurotoxicity in rodents.  The
15   immature blood-brain barrier and inefficient drug-metabolizing enzyme system make the
16   developing brain a particularly sensitive target organ to the effects of methanol exposure.

     4.4.2. Inhalation  Studies
17          A review by Carson et al. (1981) has summarized a number of older reports of studies on
18   the toxicological consequences of methanol exposure. In one example relevant to neurotoxicity,
19   the review cites a  research report of Chen-Tsi (1959) who exposed 10 albino rats/group (sex and
20   strain unstated) to 1.77 and 50 mg/m3 (1.44 and 40.7 ppm) methanol vapor, 12 hours/day, for
21   3 months. Deformation of dendrites, especially the dendrites of pyramidal cells, in the cerebral
22   cortex was included in the description of histopathological changes observed in adult animals
23   following exposure to 50 mg/m3 (40.7 ppm) methanol vapor. One out often animals exposed to
24   the lower methanol concentration also displayed this feature.
25          Information on the neurotoxicity of methanol inhalation exposure in adult monkeys
26   (M fascicularis) has come from NEDO (1987) which describes the results of a number of
27   experiments. The  study included an acute study, a chronic study monkeys,  and a repeated
28   exposure experiment (of variable duration depending upon exposure level), followed by recovery
29   period (1-6 months), and an experiment looking at chronic formaldehyde exposure (1 or 5 ppm),
30   a combustion product of methanol. This last experiment was apparently only a pilot and included
31   only one monkey  per exposure condition.
32          In the chronic experiment 8 monkeys were included per exposure level (control,  10, 100,
33   1,000 ppm or 13,  131, and 1,310 mg/m3, respectively,  for 21 hours/day); however, animals were
34   serially sacrificed at 3 time points: 7 months, 19 months, or >26 months. This design reduced
                                              4-53        DRAFT-DO NOT CITE OR QUOTE

-------
 1   the number of monkeys at each exposure level to 2 subjects at 7 months and 3 subjects at the
 2   subsequent time points (see Section 4.2.2).  One of the 3 animals receiving 100 ppm methanol
 3   and scheduled for sacrifice at 29 months was terminated at 26 months.
 4          Histopathologically, no overt degeneration of the retina, optical nerve, cerebral cortex, or
 5   other potential target organs (liver and kidney) was reported in the chronic experiment.
 6   Regarding the peripheral nervous system, 1/3  monkeys exposed to 100 ppm (131 mg/m3) and 2/3
 7   exposed to 1,000 ppm (1,310 mg/m3) for 29 months  showed slight but clear changes in the
 8   peroneal nerves. There was limited evidence  of CNS degeneration inside the nucleus of the
 9   thalamus of the brain at exposure to 100 ppm  (131 mg/m3) or 1,000 ppm (1,310 mg/m3) for
10   7 months or longer. Abnormal appearance of stellate cells  (presumed astroglia) within the
11   cerebral white matter was also observed in a high proportion (7/8 in both mid- and high-exposure
12   groups) of monkeys exposed to 100 ppm and  1,000 ppm for 7 months or more. All monkeys that
13   had degeneration of the inside nucleus of the thalamus also had degeneration of the cerebral
14   white matter. According to NEDO (1987), the stellate cell  response was transient and "not
15   characteristic of degeneration." The authors also noted that the stellate cell response was "nearly
16   absent in normal monkeys in the control group" and  "in the groups exposed to a large quantity of
17   methanol or for a long time their presence tended to become permanent, so a relation to the long
18   term over which the methanol was inhaled is suspected."  However, all control group responses
19   are reported in a single table in the section of  the NEDO (1987) report that describes the acute
20   monkey study, with no indication as to when the control group was sacrificed.
21          In the recovery experiment, monkeys were exposed to 1,000, 2,000, 3,000, or 5,000 ppm
22   methanol, followed by recovery periods of various duration. Monkeys exposed to 3,000 ppm for
23   20 days followed by a 6-month recovery  period experienced relatively severe fibrosis of
24   responsive stellate cells and lucidation of the medullary sheath. However, resolution of some of
25   the glial responses was noted in the longer duration at lower exposure levels, with no effects
26   observed on the cerebral white matter in monkeys exposed for 7 months to 1,000 ppm methanol
27   followed by a 6-month recovery period. In general, the results from the recovery experiment
28   corroborated  results observed in the chronic experiment. NEDO (1987) interpreted the lack of
29   glial effects after a 6-month recovery as an indication of a transient effect. The authors failed to
30   recognize that glial responses to neural damage do not necessarily persist following resolution of
31   neurodegeneration (Aschner and Kimmelberg, 1996).
32          The limited information available from the NEDO (1987) summary report suggests that
33   100 ppm (131 mg/m3) may be an effect level following continuous, chronic exposure to
34   methanol. However, the current report does not indicate a  robust dose response for the
35   neurodegenerative changes in the thalamus  and glial  changes  in the white matter. The number of
36   animals at each exposure level for each serial  sacrifice also limits statistical power
37   (2-3 monkeys/time point/exposure level). Confidence in this study is also weakened by the lack
38   of documentation for a concurrent control group.

                                               4-54        DRAFT-DO NOT CITE OR QUOTE

-------
 1          Weiss et al. (1996) exposed 4 cohorts of pregnant Long-Evans rats (10-12 dams/
 2    treatment group/cohort) to 0 or 4,500 ppm (0 and 5,897 mg/m3) methanol vapor (high-
 3    performance liquid chromatography [HPLC] grade), 6 hours/day, from GD6 to PND21. Pups
 4    were exposed together with the dams during the postnatal period. Average blood methanol levels
 5    in pups on PND7 and PND14 were about twice the level observed in dams.  However, methanol
 6    exposure had no effect on maternal gestational weight gain, litter size, or postnatal pup weight
 7    gain up to PND1846. Neurobehavioral tests were conducted in neonatal and adult offspring; the
 8    data generated from those tests were evaluated by repeated measures ANOVA.  Three
 9    neurobehavioral tests conducted in 13-26 neonates/group included a suckling test, conditioned
10    olfactory aversion test, and motor activity test.  In contrast to earlier test results reported by
11    Infurna and Weiss (1986), methanol exposure had no effect on suckling and olfactory aversion
12    tests conducted on PND5 and PND10, respectively. Results of motor activity tests in the
13    methanol group were inconsistent, with decreased activity on PND18 and increased activity on
14    PND25. Tests that measured motor function, operant behavior, and cognitive function were
15    conducted in 8-13 adult offspring/group.  Some small performance differences were observed
16    between control and treated adult rats in the fixed wheel running test only when findings were
17    evaluated separately by sex and cohort. The test requires the adult rats to run in a wheel and
18    rotate it a certain amount of times in order to receive a food reward. A stochastic spatial
19    discrimination test examined the  rats' ability to learn patterns of sequential responses. Methanol
20    exposure had no effect on their ability to learn the first pattern of sequential responses, but
21    methanol-treated rats  did not perform as well on the reversal test. The result indicated possible
22    subtle cognitive deficits as a result of methanol exposure. A morphological examination of
23    offspring brains conducted on PND1 and PND21  indicated that methanol exposure had no effect
24    on neuronal migration, numbers of apoptotic cells in the cortex or germinal zones, or
25    myelination. However, neural cell adhesion molecule (NCAM) 140 and NCAM 180 gene
26    expression in treated rats was reduced on PND4 but not 15 months after the last exposure.
27    NCAMs are glycoproteins required for neuron migration, axonal outgrowth, and establishing
28    mature neuronal function patterns.
29          Stanton et al. (1995) exposed 6-7 pregnant female Long-Evans rats/group to 0 or
30    15,000 ppm (0 and 19,656 mg/m3) methanol vapors (> 99.9% purity) for 7 hours/day on GD7-
31    GD19.  Mean serum methanol levels at the end of the  1st, 4th, 8th, and  12th days of exposure
32    were, 3,836, 3,764, 3,563, and 3,169 |ig/mL, respectively.  As calculated by authors,  dams
33    received an estimated methanol dose of 6,100 mg/kg-day.  A lower body weight on the first
      46 The fact that this level of exposure caused effects in the Sprague-Dawley rats of the NEDO (1987) study but did
      not cause a readily apparent maternal effect in Long-Evans rats of this study could be due to diffences in strain
      susceptibility.
                                               4-55        DRAFT-DO NOT CITE OR QUOTE

-------
 1    2 days of exposure was the only maternal effect; there was no increase in postimplantation loss.
 2    Dams were allowed to deliver and nurse litters. Parameters evaluated in pups included mortality,
 3    growth, pubertal development, and neurobehavioral function. Examinations of pups revealed
 4    that two pups from the same methanol-exposed litter were missing one eye; aberrant visually
 5    evoked potentials were observed in those pups. A modest but significant reduction in body
 6    weight gain on PND1, PND21, and PND35 was noted in pups from the methanol group.  For
 7    example, by PND35, male pups of dams exposed to methanol had a mean body weight of
 8    129 grams versus 139 grams in controls (p < 0.01).  However, postnatal mortality was unaffected
 9    by exposure to methanol. The study authors did not consider a 1.7-day delay in vaginal opening
10    in the methanol group to be an adverse effect. Preputial separation was not affected by prenatal
11    methanol exposure. Neurobehavioral status was evaluated using 8 different tests on specific
12    days up to PND160.  Tests included motor activity on PND13-PND21, PND30, and PND60,
13    olfactory learning and retention on PND18 and PND25, behavioral thermoregulation on
14    PND20-21, T-maze delayed alternation learning on PND23-PND24, acoustic startle reflex on
15    PND24, reflex modification audiometry on PND61-PND63, passive avoidance on PND73, and
16    visual evoked  potentials on PND160.  A single pup/sex/litter was examined in most tests, and
17    some animals were subjected to multiple tests. The statistical significance of neurobehavioral
18    testing was assessed by one-way ANOVA, using the litter as the statistical unit. Results of the
19    neurobehavioral testing indicated that methanol exposure had no effect on the sensory, motor, or
20    cognitive function of offspring under the conditions of the experiment. However, given the
21    comparatively small number of animals tested for each response, it is uncertain whether the
22    statistical  design had sufficient power to detect small compound-related changes.
23          NEDO (1987, unpublished report) sponsored a teratology study that included an
24    evaluation of postnatal effects in addition to standard prenatal endpoints in Sprague-Dawley rats.
25    Thirty-six pregnant females/group were exposed to 0, 200, 1,000, or 5,000 ppm (0, 262, 1,310,
26    and 6,552 mg/m3) methanol vapors (reagent grade) on GD7-GD17 for 22.7 hours/day. Statistical
27    significance of results was evaluated by t-test, Mann-Whitney U test, Fisher's exact test, and/or
28    Armitage's ^ test.
29          Postnatal effects of methanol inhalation were evaluated in the remaining 12 dams/group
30    that were permitted to deliver  and nurse their litters.  Effects were only observed in the
31    5,000 ppm.  There were no adverse effects on offspring body weight from methanol exposure.
32    However, the weights of some organs (brain, thyroid, thymus, and testes) were reduced in
33    8-week-old offspring following prenatal-only exposure to 5,000 ppm methanol. An unspecified
34    number of offspring were subjected to neurobehavioral testing or necropsy, but results were
35    incompletely reported.


                                               4-56        DRAFT-DO NOT CITE OR QUOTE

-------
 1          NEDO (1987, unpublished report) also contains an account of a two-generation
 2    reproductive study that evaluated the effects of pre- and postnatal methanol (reagent grade)
 3    exposure (20 hours/day) on reproductive and other organ systems of Sprague-Dawley rats and in
 4    particular the brain.  The FO generation (30 males and 30 females per exposure group)47 was
 5    exposed to 0, 10, 100, and 1,000 ppm (0, 13.1, 131, and 1,310 mg/m3) from 8 weeks old to the
 6    end of mating (males) or to the end of lactation period (females). The FI generation was exposed
 7    to the same concentrations from birth to the end of mating (males) or to weaning of 72 pups
 8    21  days after delivery (females).  Males and females of the F2 generation were exposed from
 9    birth to 21 days old (1 animal/sex/litter was exposed to 8 weeks of age).  NEDO (1987) noted
10    reduced brain, pituitary, and thymus weights, in the offspring of FO and FI rats exposed to
11    1,000 ppm methanol. As discussed in the report, NEDO (1987) sought to confirm the possible
12    compound-related effect of methanol on the brain by carrying out an  additional study in which
13    Sprague-Dawley rats were exposed to 0, 500,  1,000, and 2,000 ppm (0, 655, 1,310, and
14    2,620 mg/m3) methanol from the first day of gestation through the FI generation.  Brain weights
15    were measured in 10-14 offspring/sex/group at 3, 6, and 8 weeks of age. As illustrated in
16    Table 4-14, brain weights were significantly reduced in 3-week-old males and females exposed
17    to > 1,000 ppm.  At 6 and 8 weeks of age, brain weights were significantly reduced in males
18    exposed to > 1,000 ppm and females exposed to 2,000 ppm. Due to the toxicological
19    significance of this postnatal effect and the fact that it has not been measured or reported
20    elsewhere in the peer-reviewed methanol literature, the brain weight changes observed by NEDO
21    (1987) following gestational and postnatal exposures and following gestation-only exposure (in
22    the teratology study discussed above) are evaluated quantitatively and discussed in more detail in
23    Section 5 of this review.
      47 A second control group of 30 animals/sex was maintained in a separate room to "confirm that environmental conditions inside
      the chambers were not unacceptable to the animals."

                                                4-57        DRAFT-DO NOT CITE OR QUOTE

-------
          Table 4-14. Brain weights of rats exposed to methanol vapors during gestation and
          lactation
Offspring
age
3wk
6 wk
8wka
8wkb
Sex
Male
Female
Male
Female
Male
Female
Male
Female
Brain weight (g) (% control) at each exposure level
0 ppm
1.45 ±0.06
1.41 ±0.06
1.78 ±0.07
1.68 ±0.08
1.99 ±0.06
1.85 ±0.05
2.00 ±0.05
1.86 ±0.08
200 ppm
-
-
-
2.01 ±0.08
(100%)
1.91 ±0.06
(103%)
500 ppm
1.46 ±0.08
(101%)
1.41 ±0.07
(100%)
1.74 ±0.09
(98%)
1.71 ±0.08
(102%)
1.98 ±0.09
(99%)
1.83 ±0.07
(99%)
-
1,000 ppm
1.39±0.05C
(96%)
1.33±0.07d
(94%)
1.69±0.06d
(95%)
1.62 ±0.07
(96%)
1.88±0.08d
(94%)
1.80 ±0.08
(97%)
1.99 ±0.07
(100%)
1.90 ±0.08
(102%)
2,000 ppm
1.27±0.06e
(88%)
1.26±0.09e
(89%)
1.52±0.07e
(85%)
1.55±0.05e
(92%)
1.74±0.05e
(87%)
1.67±0.06e
(90%)
-
5,000 ppm
-
-
-
1.81±0.16d
(91%)
1.76 ±1.09
(95%)
       aExposed throughout gestation and FI generation.
       ''Exposed on gestational days 7-17 only.
       cp < 0.05, dp < 0.01, ep < 0.001, as calculated by the authors.
       Values are means ± S.D.
       Source: NEDO(1987).

 1          Burbacher et al. (1999a,1999b) carried out toxicokinetic, reproductive, developmental
 2   and postnatal neurological and neurobehavioral studies of methanol in M. fascicularis monkeys
 3   that were published by HEI in a two-part monograph. Some of the data were subsequently
 4   published in the open scientific literature (Burbacher et al., 2004a, 2004b). The experimental
 5   protocol featured exposure to 2 cohorts of 12 monkeys/group to low-exposure levels (relative to
 6   the previously discussed rodent studies) of 0, 200, 600, or 1,800 ppm (0, 262, 786, and 2,359
 7   mg/m3) methanol vapors (99.9% purity), 2.5 hours/day, 7 days/week, during a premating period
 8   and mating period (-180 days combined) and throughout the entire gestation period (-168 days).
 9   The monkeys were 5.5-13 years old and were a mixture of feral-born and colony-bred animals.
10   The outcome study included an evaluation of maternal reproductive performance (discussed in
11   Section 4.3.2) and tests to assess infant postnatal growth and newborn health, neurological
12   outcomes included reflexes, behavior, and development of visual, sensorimotor, cognitive, and
13   social behavioral  function. Blood methanol levels, clearance, and the appearance of formate
14   were also examined and are discussed in Section 3.2. The effects observed were in the absence of
15   appreciable increases in maternal blood formate levels.
                                               4-58
DRAFT-DO NOT CITE OR QUOTE

-------
 1          Neurobehavioral function was assessed in 8-9 infants/group during the first 9 months of
 2    life (Burbacher et al., 2004a, 1999b). Although results in 7/9 tests were negative, 2 effects were
 3    possibly related to methanol exposure.  The Visually Directed Reaching (VDR) test is a measure
 4    of sensorimotor development and assessed the infants' ability to grasp for a brightly colored
 5    object containing an applesauce-covered nipple. Beginning at 2 weeks after birth, infants were
 6    tested 5 times/day, 4 days/week.  Performance on this test, measured as age from birth at
 7    achievement of test criterion (successful object retrieval on 8/10 consecutive trials over 2 testing
 8    sessions), was reduced in all treated male infants. The times (days after birth) to achieve the
 9    criteria for the VDR test were 23.7 ±4.8 (n = 3), 32.4 ± 4.1 (n = 5), 42.7 ± 8.0 (n = 3), and 40.5 ±
10    12.5 (n = 2) days for males and 34.2 ± 1.8 (n = 5), 33.0 ± 2.9 (n = 4), 27.6 ± 2.7 (n = 5), and 40.0
11    ± 4.0 (n = 7) days for females in the control to 1,800 ppm groups, respectively.  Statistical
12    significance was obtained in the 1,800 ppm group when males and females were evaluated
13    together (p = 0.04) and in the 600 ppm (p = 0.007) for males only. However, there was no
14    significant difference between  responses and/or variances among the dose levels for males and
15    females combined (p = 0.244), for males only (p = 0.321) and for males only, excluding the high-
16    dose group (p = 0.182). Yet there was a significant dose-response trend for females only
17    (p = 0.0265). The extent to which VDR delays were due to a direct effect of methanol on
18    neurological development or a secondary effect due to the methanol-induced decrease in length
19    of pregnancy and subsequent prematurity is not clear.  Studies of reaching behavior have shown
20    that early motor development in pre-term human infants without major developmental disorders
21    differs from that of full-term infants (Fallang et al., 2003). Clinical studies have indicated that
22    the quality of reaching and grasping behavior in pre-term infants is generally less than that in
23    full-term infants (Fallang et al., 2003; Plantinga et al.,  1997). For this reason, measures of
24    human infant development generally involve adjustment of a child's "test age" if he or she had a
25    gestational age of fewer than 38 weeks, often by subtracting weeks premature from the age
26    measured from birth (Wilson and Cradock, 2004). When this type of adjustment is made to the
27    Burbacher et al. (2004a, 1999b) VDR data, the dose-response trend for males only becomes
28    worse (p = 0.448) and the dose-response trend for the females only is improved (p = 0.009),
29    though the variance in the data could not be modeled adequately. Thus, only the unadjusted
30    VDR response for females only exhibited a dose response that could be adequately modeled for
31    the purposes of this assessment (see Appendix C).
32          At 190-210 days of age, the Fagan  Test of infant intelligence was conducted. The
33    paradigm makes use of the infant's proclivity to direct more visual attention to novel stimuli
34    rather than familiar stimuli. The test measures the time infants spend looking at familiar versus
35    novel items.  Deficits in the Fagan task can qualitatively predict deficits in intelligence quotient
36    (IQ) measurements assessed in children at later ages (Fagan and Singer, 1983).  Control monkey

                                                4-59        DRAFT-DO NOT CITE OR QUOTE

-------
 1    infants in the Burbacher et al. (2004a, 1999b) study spent more than 62% ± 4% (mean for both
 2    cohorts) of their time looking at novel versus familiar monkey faces, while none of the treated
 3    monkeys displayed a preference for the novel faces (59% ± 2%, 54% ± 2% and 59% ± 2% in
 4    200, 600 and 1,800 ppm groups, respectively).  Unlike the VDR results discussed previously,
 5    results of this test did not appear to be gender specific and were neither statistically significant
 6    (ANOVA p = 0.38) nor related to exposure concentration. The findings indicated a cohort effect
 7    which appeared to reduce the statistical power of this analysis. The authors' exploratory analysis
 8    of differences in outcomes between the 2 cohorts indicated an effect of exposure in the second
 9    cohort and not the first cohort due to higher mean performance in controls of cohort 2 (70% +
10    5% versus 55% ± 4% for cohort 1). In addition, this latter finding could reflect the inherent
11    constraints of this endpoint.  If the control group performs at the 60% level and the most
12    impaired subjects perform at approximately the 50% chance level (worse than chance
13    performance would not be expected), the range over which a concentration-response relationship
14    can be expressed is limited.  Because of the longer latency between assessment and birth, these
15    results would not be confounded with the postulated methanol-induced decrease in gestation
16    length of the exposed groups of this study.  Negative results were obtained for the remaining
17    seven tests that evaluated early reflexes, gross motor development,  spatial and concept learning
18    and memory, and social behavior.  Infant growth and tooth eruption were  unaffected by methanol
19    exposure.

      4.4.3. Studies Employing In Vitro, S.C. and I.P. Exposures
20          There is some experimental evidence that the presence of methanol can affect the activity
21    of acetylcholinesterase (Tsakiris, et al., 2006). Although these experiments were carried out  on
22    erythrocyte membranes in vitro, the apparent compound-related changes may have implications
23    for possible impacts of methanol and/or its metabolites on acetylcholinesterase at other centers,
24    such as the brain.  Tsakiris et al. (2006) prepared erythrocyte ghosts from blood samples of
25    healthy human volunteers by repeated freezing-thawing. The ghosts were incubated for 1 hour at
26    37°C in 0, 0.07, 0.14, 0.6 or  0.8 mmol/L methanol and the specific  activities of
27    acetylcholinesterase monitored. Respective values  (in change of optical  density units/minute-mg
28    protein) were 3.11 ± 0.15, 2.90 ± 0.10, 2.41 ± 0.10 (p < 0.05), 2.05  ± 0.11 (p < 0.01), and 1.81 ±
29    0.09 (p < 0.001). More recently, Simintzi et al.  (2007) carried out an in vitro experiment to
30    investigate the effects of aspartame metabolites, including methanol, on  1) a pure preparation of
31    acetylcholinesterase, and 2)  the same activity in homogenates of frontal cortex prepared from the
32    brains of (both sexes of) Wistar rats.  The activities  were measured  after  incubations with 0, 0.14,
33    0.60, or 0.8 mmoles/L (0, 4.5,  19.2, and 25.6 mg/L) methanol,  and  with methanol mixed with the
34    other components of aspartame metabolism, phenylalanine and aspartic acid. After incubation at
                                               4-60        DRAFT-DO  NOT CITE OR QUOTE

-------
 1    37°C for 1 hour, the activity of acetylcholinesterase was measured spectrophotometrically.  As
 2    shown in Table 4-15, the activities of the acetylcholinesterase preparations were reduced dose
 3    dependency after incubation in methanol. Similar results were also obtained with the other
 4    aspartame metabolites, aspartic acid, and phenylalanine, both individually or as a mixture with
 5    methanol. While the implications of this result to the acute neurotoxicity of methanol are
 6    uncertain, the authors speculated that methanol may bring about these changes through either
 7    interactions with the lipids of rat frontal cortex or perturbation of proteinaceous components.

          Table 4-15. Effect of methanol on Wistar rat acetylcholinesterase activities
Methanol concentration
(mmol/L)
Control
0.14
0.60
0.80
Acetylcholinesterase activity (AOD/min-mg)
Frontal cortex
0.269 ±0.010
0.234 ±0.007a
0.223 ± 0.009b
0.204 ± 0.008b
Pure enzyme
1.23 ±0.04
1.18 ±0.06
1.05±0.04b
0.98±0.05b
      Values are means ± S.D. for four experiments. The average value of each experiment was derived from three
      determinations of each enzyme activity.
      ><0.01.
      V< 0.001.
      Source: Simintzi et al. (2007).

 8           In another experiment of relevance to neurotoxicity, the impact of repeat methanol
 9    exposure on amino acid and neurotransmitter expression in the retina, optic nerve, and brain was
10    examined by Gonzalez-Quevedo et al. (2002).  The goal of the study was to determine whether a
11    sustained increase in formate levels, at concentrations below those known to produce toxic
12    effects from acute exposures, can induce biochemical changes in the retina, optical nerve, or
13    certain regions of the brain. Male Sprague-Dawley rats (5-7/group; 100-150 g) were divided
14    into 6 groups and treated for 4 weeks according to the following plan.  Four groups of animals
15    received tap water ad libitum as drinking water for 1 week. During the second week, groups 1
16    and 2 (control and methanol respectively) received saline subcutaneously, (s.c.) and groups 3 and
17    4 (methotrexate48 [MTX] and methotrexate-methanol [MTX-methanol], respectively) received
18    MTX s.c. (0.2 mg/kg-day).  During the 3rd week, MTX was reduced to 0.1 mg/kg and 20%
19    methanol (2g/kg-day) was given i.p. to groups  2 (methanol) and 4 (MTX-methanol). Groups 1
20    (control) and 3 (MTX)  received equivalent volumes of saline administered i.p.  The treatment
21    was continued until the end of the fourth week. Groups 5, (taurine49 [Tau]) and 6, (Tau-MTX-
      48 Methotrexate depletes folate stores (resulting in an increase in the formate levels of methanol exposed animals) by
      interfering with tetrahydrofolate(THF) regeneration (Dorman et al., 1994).
      49 Taurine plays and important role in the CNS, especially in the retina and optical nerve, and was administered here
      to explore its possible protective effect (Gonzalez-Quevedo et al., 2002)
                                                 4-61        DRAFT-DO NOT CITE OR QUOTE

-------
 1    methanol) received 2% Tau in their drinking water ad libitum during the first 4 weeks, after
 2    which they were treated in the same manner as groups 1 and 4, respectively. Weights were
 3    documented weekly on all animals. Blood for formate and amino acid determinations and
 4    biopsy samples of retina, optic nerve, hippocampus, and posterior cortex of each animal were
 5    collected at the end of the experiment.  Formate levels were not affected by Tau alone or MTX
 6    alone. While methanol alone increased blood formate levels, MTX-methanol, and Tau-MTX-
 7    methanol produced a threefold increase in blood formate levels as compared to controls and a
 8    twofold increase as compared to methanol alone. The amino acids aspartate, glutamate,
 9    asparagine, serine, histidine, glutamine, threonine, glycine, arginine, alanine, hypotaurine,
10    gamma-aminobutyric acid (which is also a neurotransmitter), and tyrosine were measured in
11    blood, brain, and retinal regions.
12          None of the amino acids measured were altered in the blood of methanol-, MTX-,  or
13    MTX-methanol-treated animals.  Tau was increased in the blood of animals treated with taurine
14    in the drinking water (Tau and Tau-MTX-methanol) and histidine was increased in the Tau group
15    but not in the Tau-MTX-methanol group.
16          The levels of aspartate, Tau, glutamine, and glutamate were found to be altered by
17    treatment in various areas of the brain.  Aspartate was increased in the optic nerve of animals
18    treated with MTX-methanol and Tau-MTX-methanol, indicating a possible relation to formate
19    accumulation. The authors note that L-aspartate is a major excitatory amino acid in the brain and
20    that increased  levels of excitatory amino acids can trigger neuronal cell damage and death (Albin
21    and Greenamyre, 1992). Aspartate, glutamine and Tau were found to be increased with respect
22    to controls in the hippocampus of the three groups receiving methanol.  Glutamate was
23    significantly increased in the hippocampus in  the methanol and the Tau-MTX-methanol groups
24    with respect to controls, but no statistically significant difference was found in the MTX-
25    methanol group when compared to controls, methanol alone, or the Tau-MTX-methanol groups.
26    The authors suggest that increased levels of aspartate  and glutamine in the hippocampus could
27    provide an explanation  for some of the CNS symptoms observed in methanol poisonings on the
28    basis of their observed impact on cerebral arteries (Huang et al., 1994).  The fact that these
29    increases resulted primarily from methanol without MTX is significant in that it indicates
30    methanol can cause excitotoxic effects without formate mediation. The treatments used did not
31    produce any significant changes in amino acid levels in the posterior cortex.
32          The neurotransmitters serotonin (5-HT) and dopamine (DA) and their respective
33    metabolites, 5-hydroxyindolacetic acid (5-HIAA) and dihydroxyphenylacetic acid (DOPAC),
34    were measured in the brain regions described. The levels of these monoamines were not affected
35    by formate accumulation, as the only increases were observed for 5-HT and 5-HIAA following
36    methanol-only exposure.  5-HT was increased in the retina and hippocampus of methanol-only

                                               4-62       DRAFT-DO NOT CITE OR QUOTE

-------
 1    treated animals, and the metabolite 5-HIAA was increased in the hippocampus of methanol-only
 2    treated animals; DA and DOPAC levels were not altered by the treatments in any of the areas
 3    measured.  The posterior cortex did not show any changes in monoamine levels for any treatment
 4    group.
 5          Rajamani et al. (2006) examined several oxidative stress parameters in male Wistar rats
 6    following methotrexate-induced folate deficiency. Animals (6/group) were divided into 3e
 7    groups: saline controls, methotrexate (MTX) controls, and MTX-methanol treated animals.
 8    Animals in the MTX-only group were treated with 0.2 mg/kg-day MTX s.c. injection for 7 days
 9    and following confirmation of folate deficiency, received either saline for MTX control and
10    saline controls or a single dose of 3 g/kg methanol (20% w/v in saline) i.p. on day  8. On the 9th
11    day, all animals were sacrificed and blood and tissue samples were collected.  The optic nerve,
12    retina, and brain were collected and the brain was dissected into the following regions: cerebral
13    cortex, cerebellum, mid-brain, pons medulla, hippocampus and hypothalamus. Each region was
14    homogenized, then centrifuged at 300 x g for 15 minutes and the supernatant was examined for
15    indicators of oxidative stress including the free radical scavengers superoxide dismutase (SOD),
16    CAT, glutathione peroxidase, and reduced GSH levels.  The levels of protein thiols, protein
17    carbonyls, and amount of lipid peroxidation were also measured. Compared to controls the
18    levels of SOD, CAT, GSH peroxidase, oxidized GSH, protein carbonyls and lipid peroxidation
19    were elevated in all of the brain regions where it was measured, with greater increases observed
20    in the MTX-methanol treated animals than in the MTX alone group. The level of GSH and
21    protein thiols was decreased in all regions of the brain, with a greater decrease observed in the
22    MTX-methanol-treated animals than MTX-treated animals.  In addition, expression of HSP70, a
23    biomarker of cellular stress, was increased in the hippocampus. Overall, these results suggest that
24    methanol treatment of folate-deficient rats results in increased oxidative stress in the brain, retina
25    and optic nerve.
26          To determine the effects of methanol intoxication on the HPA axis, a combination of
27    oxidative stress, immune and neurobehavioral parameters were observed (Parthasarathy et al.,
28    2006a). Adult male Wistar albino rats (6 animals/group) were treated with either 0 or 2.37g/kg-
29    day methanol  i.p. for 1, 15 or 30 days. Oxidative stress parameters examined included SOD,
30    CAT, GSH peroxidase, GSH, and ascorbic acid (Vitamin C).  Plasma corticosterone levels were
31    measured, and lipid peroxidation was measured in the hypothalamus and the adrenal gland. An
32    assay for DNA fragmentation was conducted in tissue from the hypothalamus, the adrenal gland
33    and the spleen. Immune function tests conducted included the footpad thickness test for delayed
34    type hypersensitivity (DTH), a leukocyte migration inhibition assay, the hemagglutination assay
35    (measuring antibody titer), the neutrophil adherence test, phagocytosis index,  and a nitroblue
36    tetrazolium (NET) reduction and adherence assay used to measure the killing ability of

                                               4-63        DRAFT-DO NOT CITE OR QUOTE

-------
 1    polymorphonuclear leukocytes (PMNs).  The open field behavior test was used to measure
 2    general locomotor and explorative activity during methanol treatment in the 30-day treatment
 3    group, with tests conducted on days 1, 4, 8, 12, 16, 20, 24, and 28. All enzymatic (SOD, CAT,
 4    and GSH peroxidase) and nonenzymatic antioxidants (GSH and Vitamin C) were significantly
 5    increased in the 1-day methanol-exposed group as compared to controls.  However, with
 6    increasing time of treatment, all of the measured parameters were significantly decreased when
 7    compared with control animals. Lipid peroxidation was significantly increased in both the
 8    hypothalamus and the adrenal gland at 1, 15, and 30 days, with the 30-day treated animals also
 9    significantly increased when compared to the 15-day methanol-treated animals.
10          Leukocyte migration and antibody titer were both  significantly increased over controls
11    for all time points, while footpad thickness was significantly decreased in 15- and 30-day treated
12    animals. Neutrophil adherence was significantly decreased after 1 and 30 days of exposure. A
13    significant decrease in the NET reduction and adherence was found when comparing PMNs from
14    the 30-day treated animals with cells from the 15-day methanol-treated group.
15          The open field behavior tests showed a significant decrease in ambulation from the 4th
16    day on and significant decreases in rearing and grooming from the 20th day on. A significant
17    increase was observed in immobilization from the 8th day on and in fecal bolus from the 24th
18    day on in methanol-exposed animals.
19          While corticosterone levels were significantly increased following 1 or 15 days of
20    methanol treatment, they were significantly decreased after 30 days of treatment, as compared to
21    controls. Following 30 days of methanol treatment, DNAfrom the hypothalamus, the adrenal
22    gland, and the spleen showed significant fragmentation. The authors conclude that exposure to
23    methanol-induced oxidative stress, disturbs HPA-axis function, altering corticosterone levels and
24    producing effects in several nonspecific and specific immune responses.

      4.5. IMMUNOTOXICITY
25          Parthasarathy et al. (2005a) provided data on the impact of methanol on neutrophil
26    function in an experiment in which 6 male Wistar rats/group were given a single i.p. exposure of
27    2,370 mg/kg methanol mixed 1:1 in saline. Another group of 6 animals provided blood samples
28    that were incubated with methanol in vitro at a methanol concentration equal to that observed in
29    the in vivo-treated animals 30 and 60 minutes  postexposure.  Total and differential leukocyte
30    counts were measured from these groups in comparison to in vivo and in vitro controls.
31    Neutrophil adhesion was determined by comparing the neutrophil index in the untreated blood
32    samples to those that had been passed down a nylon fiber  column. The cells' phagocytic ability
33    was evaluated by their ability to take up heat-killed Candida albicans. In another experiment,
34    neutrophils were assessed for their killing potential by measuring their ability to take up then
                                                4-64        DRAFT-DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
convert NET to formazan crystals.50  One hundred neutrophils/slides were counted for their total
and relative percent formazan-positive cells.
       The blood methanol concentrations 30 and 60 minutes after dosing were 2,356 ±162 and
2,233 ±146 mg/L, respectively.  The mean of these values was taken as the target concentration
for the in vitro methanol incubation.  In the in vitro studies, there were no differences in total and
differential leukocyte counts, suggesting that no lysis of the cells had occurred at this methanol
concentration. This finding contrasts with the marked difference in total leukocytes observed as
a result of methanol incubation in vivo, in which, at 60 minutes after exposure, 16,000 ± 1,516
cells/mm3 were observed versus 23,350 ± 941 in controls (p < 0.001). Some differences in
neutrophil function were observed in blood samples treated with methanol in vitro and in vivo.
These differences are illustrated for the 60-minute postexposure samples in Table 4-16.

     Table 4-16. Effect of methanol on neutrophil functions in in vitro and in vivo studies in
     male Wistar rats
Parameter
Phagocytic index (%)
Avidity index
NET reduction (%)
Adherence (%)
In vitro studies (60 minutes)
Control
89.8 ±3.07
4.53 ±0.6
31.6 ±4.6
50.2 ±5.1
Methanol
81.6±2.2a
4.47 ±0.7
48.6±4.3b
39.8±2.4a
In vivo studies (60 minutes)
Control
66.0 ±4.8
2.4 ±0.1
4.6 ±1.2
49.0 ±4.8
Methanol
84.0 ± 7.0b
3.4±0.3a
27.0 ± 4.6b
34.6±4.0b
12
13
14
15
16
17
18
19
20
21
22
23
Values are means ± S.D. for six animals.
><0.01.
V< 0.001.
Source: Parthasarathy et al. (2005a).

       Parthasarathy et al. (2005a) observed differences in the neutrophil functions of cells
exposed to methanol in vitro versus in vivo, most notably in the phagocytic index that was
reduced in vitro but significantly increased in vivo. However, functions such as adherence and
NET reduction showed consistency in the in vitro and in vivo responses. The authors noted that,
by and large, the in vivo effects of methanol  on neutrophil function were more marked than those
in cells exposed in vitro.
       Another study by Parthasarathy et al. (2005b) also exposed 6 male Wistar rats/group i.p.
to methanol at approximately 1/4 the LD50 (2.4 g/kg). The goal was to further monitor possible
methanol-induced alterations in the activity of isolated neutrophils and other immunological
parameters. The exposure protocol featured  daily injections of methanol for up to 30 days in the
presence or absence of sheep RBCs.  Blood samples were assessed for total and differential
leukocytes, and isolated neutrophils were monitored for changes in phagocytic and avidity
      50 Absence of NET reduction indicates a defect in some of the metabolic pathways involved in intracellular
      microbial killing.
                                                4-65
                                                       DRAFT-DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
indices, NET reduction, and adherence. In the latter test, blood samples were incubated on a
nylon fiber column, then eluted from the column and rechecked for total and differential
leukocytes. Phagocytosis was monitored by incubating isolated buffy coats from the blood
samples with heat-killed C. albicam.  NET reduction capacity examined the conversion of the
dye to formazan crystals within the cytoplasm. The relative percentage of formazan-positive
cells in each blood specimen gave a measure of methanol's capacity to bring about cell death.
As tabulated by the authors, there was a dose-dependent reduction in lymphoid organ weights
(spleen, thymus, and lymph node) in rats exposed to methanol for 15 and 30 days via i.p.
injection, irrespective of the presence of sheep RBCs.  Methanol also appeared to result in a
reduction in the total or differential neutrophil count.  These and potentially related changes to
neutrophil function are shown in Table 4-17.

     Table 4-17. Effect of intraperitoneally injected methanol  on total and differential
     leukocyte counts and neutrophil function tests in male Wistar rats
Parameter
Without sheep red blood cell treatment
Control
15-day
methanol
30-day
methanol
With sheep red blood cell treatment
Control
15-day
methanol
30-day
methanol
Organ weights (mg)
Spleen
Thymus
Lymph node
1223 ± 54
232 ±12
32 ±2
910±63a
171 ±7a
24±3a
696 ± 83a'b
121 ± 10a'b
16 ± 2a'b
1381 ±27
260 ±9
39 ±2
1032 ±39a
172 ± 10a
28±la
839 ± 35a'b
130 ± 24a'b
23 ± la'b
Leukocyte counts
Total leukocytes
% neutrophils
% Lymphocytes
23,367 ± 946
24 ±8
71±7
16,592 ±1219a
21 ±3
76 ±3
13,283 ±
2553a'b
16±3a
79 ±5
18,633 ± 2057
8±3
89 ±4
16,675 ± 1908
23±4a
78.5 ±4a
14,067 ± 930a'b
15 ± 5a'b
82 ±6
Neutrophil function tests
Phagocytic index
(%)
Avidity index
NET reduction (%)
Adherence (%)
91.0 ±2.0
2.6 ±0.3
6.3 ±2.0
49.0 ±5.0
80.0 ± 4.0a
3.2±0.5a
18.2±2.0a
44.0 ±5.0
79.0±2.0a
3.2±0.1a
15.0±1.0a'b
29.5 ± 5.0a'b
87.0 ±4.0
4.1±0.1
32.0 ±3.3
78.0 ±9.2
68.0±3.0a
2.6±0.3a
22.0±3.0a
52.0±9.0a
63.0±4.0a
2.1±0.3a
19.0±2.4a
30.0±4.3a'b
     Values are means ± S.D. (n = 6).
     ap < 0.05 from respective control.
     bp < 0.05 between 15-and 30-day treatment groups.
     Source: Parthasarathy et al. (2005b).

12          The study provided data that showed altered neutrophil functions following repeated
13   daily exposures of rats to methanol for periods up to 30 days. This finding is indicative of a
14   possible effect of methanol on the immunocompetence of an exposed host.
                                                4-66
                                                      DRAFT-DO NOT CITE OR QUOTE

-------
 1          Parthasarathy et al. (2006a) reported on additional immune system indicators as part of a
 2    study to determine the effects of methanol intoxication on the HPA axis. As described in
 3    Section 4.4.3. immune function tests conducted included the footpad thickness test for DTK, a
 4    leukocyte migration inhibition assay, the hemagglutination assay (measuring antibody titer), the
 5    neutrophil adherence test, phagocytosis index, and a NET reduction and adherence assay used to
 6    measure the killing ability of PMNs.
 7          Leukocyte migration and antibody titer were both significantly increased over controls
 8    for all time points, while footpad thickness was significantly deceased in 15- and 30-day treated
 9    animals.  Neutrophil adherence was significantly decreased after 1 and 30  days of exposure. A
10    significant decrease in the NET reduction and adherence was found when  comparing PMNs from
11    the 30-day treated animals with cells from the 15-day methanol-treated group.
12          Parthasarathy et al. (2007) reported the effects of methanol on a number of specific
13    immune functions. As before, 6 male Wistar rats/group were treated with 2,370 mg/kg methanol
14    in a 1:1 mixture in saline administered intraperitoneally for 15 or 30 days.  Animals
15    scheduled/designated for termination on day 15 were immunized intraperitoneally with 5 x 109
16    sheep RBCs on the 10th day. Animals scheduled for day 30 termination were immunized on the
17    25th day.  Controls were animals that were not exposed to methanol but immunized with sheep
18    RBCs as described above. Blood samples were obtained from all animals  at sacrifice and
19    lymphoid organs including the adrenals, spleen, thymus, lymph nodes, and bone marrow were
20    removed.  Cell suspensions were counted and adjusted to 1  x 108 cells/mL. Cell-mediated
21    immune responses were assessed using a footpad thickness assay and a leucocyte migration
22    inhibition (LMI) test, while humoral immune responses were determined by a hemagglutination
23    assay, and by monitoring cell counts in spleen, thymus, lymph nodes, femoral bone marrow, and
24    in splenic lymphocyte subsets.  Plasma levels of corticosterone were measured along with levels
25    of such cytokines as TNF-a, IFN-y, IL-2, and IL-4. DNA damage in splenocytes and thymocytes
26    was also monitored using the Comet assay.
27          Table 4-18  shows  decreases in the animal weight/organ weight ratios for spleen, thymus,
28    lymph nodes and adrenal gland as a result of methanol exposure. However, the splenocyte,
29    thymocyte, lymph node, and bone marrow cell counts were time-dependently lower in methanol-
30    treated animals.
                                               4-67       DRAFT-DO NOT CITE OR QUOTE

-------
          Table 4-18. Effect of methanol exposure on animal weight/organ weight ratios and on
          cell counts in primary and secondary lymphoid organs of male Wistar rats.
Organ
Immunized
Control
15 days
30 days
Animal weight/organ weight ratio
Spleen
Thymus
Lymph node
Adrenal
3.88 ±0.55
1.35 ±0.29
0.10 ±0.01
0.14 ±0.01
2.85±0.36a
0.61±0.06a
0.08±0.01a
0.15 ±0.01
2.58±0.45a
0.63 ±0.04a
0.06 ± 0.02a
0.12±0.01a'b
Cell counts
Splenocytes (x 108)
Thymocytes (x 108)
Lymph node (x 107)
Bone marrow (x 107)
5.08 ±0.06
2.66 ±0.09
3.03 ±0.04
4.67 ±0.03
3.65±0.07a
1.95±0.03a
2.77 ± 0.07a
3.04±0.09a
3.71±0.06a
1.86±0.09a
2.20±0.06a'b
2.11±0.05a'b
     Values are means ± six animals.
     ap < 0.05 versus control groups.
     bp < 0.05 versus 15-day treated group.
     Source: Parthasarathy et al. (2007).

 1          Parthasarathy et al. (2007) also documented their results on the cell-mediated and
 2   humoral immunity induced by methanol.  Leucocyte migration was significantly increased
 3   compared to control animals, an LMI of 0.82 ± 0.06 being reported in rats exposed to methanol
 4   for 30 days. This compares to an LMI of 0.73 ± 0.02 in rats exposed for 15 days and 0.41 ± 0.10
 5   in controls.  By contrast, footpad thickness and antibody titer were decreased significantly in
 6   methanol-exposed animals compared to controls (18.32 ± 1.08,  19.73 ± 1.24, and 26.24 ± 1.68%
 7   for footpad thickness; and 6.66 ± 1.21, 6.83 ± 0.40, and 10.83 ± 0.40 for antibody titer in 30-day,
 8   15-day exposed rats, and controls, respectively).
 9          Parthasarathy et al. (2007) also provided data in a histogram that showed a significant
10   decrease in the absolute numbers of Pan T cells, CD4, macrophage, major histocompatibility
11   complex (MHC) class II molecule expressing cells, and B cells of the methanol-treated group
12   compared to controls. The numbers of CDS cells were unaffected. Additionally, as illustrated in
13   the report, DNA single strand breakage was increased in immunized splenocytes and thymocytes
14   exposed to methanol versus controls. Although some fluctuations were seen in corticosterone
15   levels, the apparently statistically significant  change versus controls in 15-day exposed rats was
16   offset by a decrease in 30-day exposed animals. Parthasarathy et al. (2007) also tabulated the
17   impacts of methanol exposure on cytokine levels; these values are shown in Table 4-19.
                                               4-68
DRAFT-DO NOT CITE OR QUOTE

-------
          Table 4-19. The effect of methanol on serum cytokine levels in male Wistar rats
Cytokines (pg/mL)
IL-2
IL-4
TNF-a
IFN-y
Immunized
Control
1810 ±63.2
44.8 ±2.0
975 ± 32.7
1380 ±55.1
15 days
1303.3 ±57.1a
74.0 ± 5. la
578.3 ± 42.6a
9616±72.7a
30 days
1088.3 ± 68.8a'b
78.8±4.4a
585 ± 45a
950±59.6a
      Values are means ± six animals.
      ap < 0.05 versus control groups.
      bp < 0.05 versus 15-day treated group.
      Source: Parthasarathy et al. (2007).

 1          Drawing on the results of DNA single strand breakage in this experiment, the authors
 2    speculated that methanol-induced apoptosis could suppress specific immune functions such as
 3    those examined in this research report. Methanol appeared to suppress both humoral and cell-
 4    mediated immune responses in exposed Wistar rats.

      4.6. MECHANISTIC DATA AND OTHER STUDIES IN SUPPORT OF THE MOA
 5          While the role of the methanol metabolite, formate, in inducing the toxic consequences of
 6    acute exposure to methanol, including ocular toxicity and metabolic acidosis, is well established
 7    in humans (see Section 4.1), there is controversy over the possible roles of the parent compound,
 8    metabolites, and folate deficiency (potentially associated with methanol metabolism) in the
 9    developmental neurotoxicity of methanol.  Experiments that have attempted to address these
10    issues are reviewed in the following paragraphs.

      4.6.1. Role of Methanol and Metabolites in the Developmental Toxicity of Methanol
11          Dorman et al. (1995) conducted a series of in vitro and in vivo studies that provide
12    information for identifying the proximate teratogen associated with developmental toxicity in
13    CD-I mice. The studies used CD-I ICR BR (CD-I) mice, HPLC grade methanol, and
14    appropriate controls. PK and developmental toxicity parameters were measured in mice exposed
15    to sodium formate (750 mg/kg by gavage), a 6-hour methanol inhalation (10,000 or 15,000 ppm),
16    or methanol gavage (1.5 g/kg). In the in vivo inhalation study, 12-14 dams/ group were exposed
17    to 10,000 ppm methanol for 6 hours on GD8,51 with and without the administration of
18    fomepizole (4-methylpyrazole) to inhibit the metabolism of methanol by ADH1. Dams were
19    sacrificed on GD10, and fetuses were examined for neural tube patency. As shown in
20    Table 4-20, the incidence of fetuses with open neural tubes was significantly increased in the
      51 Dorman et al. (1995) state that GD8 was chosen because it encompasses the period of murine neurulation and the
      time of greatest vulnerability to methanol-induced neural tube defects.
                                               4-69
DRAFT-DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
methanol group (9.65% in treated versus 0 in control) and numerically but not significantly
increased in the group treated with methanol and fomepizole (7.21% in treated versus 0 in
controls).  These data should not be interpreted to suggest that a decrease in methanol
metabolism is protective. As discussed in Section 3.1, rodents metabolize methanol via both
ADH1 and CAT. This fact and the Dorman et al. (1995) observation that maternal formate levels
in blood and decidual swellings (swelling of the uterine lining) did not differ in dams exposed to
methanol alone or methanol and fomepizole suggest that the role of ADH1 relative to CAT and
nonenzymatic methanol clearance is not of great significance in adult rodents.

    Table 4-20. Developmental outcome on GD10 following a 6-hour 10,000 ppm
    (13,104 mg/m3) methanol inhalation by CD-mice or formate gavage (750 mg/kg) on
    GD8
Treatment
Air
Air/fomepizole
Methanol
Methanol/fomepizole
Water
Formate
No. of litters
14
14
12
12
10
14
Open neural tubes (%)
2.29 ±1.01
2.69 ±1.19
9.65±3.13a
7.21 ±2.65
0
2.02 ±1.08
Head length (mm)
3. 15 ±0.03
3.20 ±0.05
3.05 ±0.07
3.01 ±0.05
3.01 ±0.07
2.91 ±0.08
Body length (mm)
5.89 ±0.07
5.95 ±0.09
5.69 ±0.13
5.61 ±0.11
5.64 ±0.11
5.49 ±0.12
 9
10
11
12
13
14
15
16
17
18
19
20
21
22
Values are means ± S.D.
a/> < 0.05, as calculated by the authors.
Source: Dorman etal. (1995) (adapted).

       The data in Table 4-20 suggest that the formate metabolite is not responsible for the
observed increase in open neural tubes in CD-I mice following methanol exposure.  Formate
administered by gavage (750 mg/kg) did not increase this effect despite the fact that this formate
dose produced the same toxicokinetic profile as a 6-hour exposure to 10,000 ppm methanol
vapors (1.05 mM formate in maternal blood and 2.0 mmol formate/kg in decidual swellings).
However, the data are consistent with the hypotheses that the formaldehyde metabolite of
methanol may play a role. Both CAT and ADH1 activity are immature at days past conception
(DPC)8 (Table 4-21). If fetal ADH1 is more mature than fetal CAT, it is conceivable that the
decrease in the open neural tube response observed for methanol combined with fomepizole
(Table 4-20) may be due to fomepizole having a greater effect on the metabolism of fetal
methanol to formaldehyde than is observed in adult rats. Unfortunately, the toxicity studies were
carried out during a period of development where ADH1 expression and activity are just starting
to develop (Table 4-21); therefore, it is uncertain whether any ADH1 was present in the fetus to
be inhibited by fomepizole.
                                               4-70
                                                     DRAFT-DO NOT CITE OR QUOTE

-------
          Table 4-21. Summary of ontogeny of relevant enzymes in CD-I mice and humans



Somites
CAT
mRNA
activity3
embryo
VYS
ADH1
mRNA
activity
embryo
VYS
ADH3
mRNA
activity
embryo
VYS
CD-I Mouse
Days Past Conception (DPC)
6.5







-




+



7.5







-




+




(8-12)



1
10



320
240



300
500
8.5







-




+




(13-20)



10
15



460
280



490
500
9.5
(21-29)



20
20

+

450
290

+

550
550
Human
Trimesters
1

N/A





+




—



2

N/A





+




—



o
6

N/A





+




+



     ""Activity of CAT and ADH1 are expressed as nmol/minute/mg and pmol/minute/mg, respectively.
     Source: Harris et al. (2003).

 1          Dorman et al. (1995) provide additional support for their hypothesis that methanol's
 2   developmental effects in CD-I mice are not caused by formate in an in vitro study involving the
 3   incubation of GD8 whole CD-I mouse embryos with increasing concentrations of methanol or
 4   formate.  Developmental anomalies were observed on GD9, including cephalic dysraphism,
 5   asymmetry and hypoplasia of the prosencephalon, reductions of brachial arches I and II,
 6   scoliosis, vesicles on the walls of the mesencephalon, and hydropericardium (Table 4-22).  The
 7   concentrations of methanol used for embryo incubation (0-375  mM) were chosen to be broadly
 8   equivalent to the peak methanol levels in plasma that have been observed (approximately
 9   100 mM) after a single 6-hour inhalation exposure to 10,000 ppm (13,104 mg/m3). As discussed
10   above, these exposure conditions induced an increased incidence of open neural tubes on GD10
11   embryos when pregnant female CD-I mice were exposed on GD8. (Table 4-20).  Embryonic
12   lesions such as cephalic dysraphism, prosencephalic lesions, and brachial arch hypoplasia were
13   observed with 250 mM (8,000 mg/L) methanol and 40 mM (1,840 mg/L) formate.  The study
14   authors noted that a formate concentration of 40 mM (1,840 mg/L) greatly exceeds blood
15   formate levels in mice inhaling 15,000 ppm methanol (0.75 mM = 35 mg/L), a teratogenic dose.
                                              4-71
DRAFT-DO NOT CITE OR QUOTE

-------
          Table 4-22. Dysmorphogenic effect of methanol and formate in neurulating CD-I
          mouse embryos in culture (GD8)
Treatment
Vehicle
Methanol
Formate
Concentration
(mM)

62
125
187
250
375
4
8
12
20
40
Live embryos
Total
20
13
14
13
15
12
12
13
9
16
16
No.
abnormal
3
1
5
7
7
7
2
5
5
7
14a
Cephalic dysraphism
Severe
0
0
1
2
2
6a
0
1
0
2
10a
Moderate
2
0
0
4
5
5
0
5
5
5
4
Total
2
0
2
6
7
lla
0
6
5
7
14a
Prosencephalic lesions
Hypoplasia
2
1
2
3
7a
9a
2
4
1
2
3
Asymmetry
0
0
2
1
1
1
0
2
2
1
5a
Total
2
1
4
4
8
10a
2
6
3
3
8
Brachial
arch
hypoplasia
0
0
1
1
6a
8a
1
0
0
1
13a
      a/> < 0.05, as calculated by the authors.
      Source: Dormanetal. (1995) (adapted).

 1          As discussed in Section 4.3.3, a series of studies by Harris et al. (2004, 2003) also
 2    provide evidence as to the moieties that may be responsible for methanol-induced developmental
 3    toxicity.  Harris et al. (2004) have shown that among methanol and its metabolites, viability of
 4    cultured rodent embryos is most affected by formate.  In contrast, teratogenic endpoints (of
 5    interest to this risk assessment) in cultured rodent embryos are more sensitive to methanol and
 6    formaldehyde than formate. Data from these studies indicate that developmental toxicity may be
 7    more related to formaldehyde than methanol, as formaldehyde-induced teratogenicity occurs at
 8    several orders of magnitude lower than methanol (Table 4-14) (Hansen et al., 2005; Harris et al.,
 9    2004). It should also be noted that CAT, ADH1, and ADH3 activities are present in both the rat
10    embryo and VYS at stages as early as 6-12 somites (Harris et al., 2003); thus, it is presumable
11    that in these ex vivo studies methanol is metabolized to formaldehyde and formaldehyde is
12    subsequently metabolized to S-formylglutathione.
13          Studies involving GSH also lend support that formaldehyde may be a key proximal
14    teratogen. Inhibition of GSH synthesis with butathione sulfoximine (BSO) has little effect on
15    developmental toxicity endpoints, yet treatment with BSO and methanol or formaldehyde
16    increases developmental toxicity (Harris et al., 2004). Among the enzymes involved in methanol
17    clearance, only ADH3-mediated metabolism of formaldehyde is GSH dependent.  This
18    hypothesis that ADH3-mediated metabolism of formaldehyde is important for the amelioration
19    of methanol's developmental toxicity is also supported by the diminished ADH3 activity in the
20    mouse versus rat embryos, which is consistent with the greater sensitivity of the mouse to
                                               4-72       DRAFT-DO NOT CITE OR QUOTE

-------
 1    methanol developmental toxicity (Harris et al., 2003) (Section 4.3.3).  Similarly reasonable
 2    explanations for this greater mouse sensitivity are not readily apparent for the two MO As
 3    described below that attribute methanol toxicity to methanol metabolism per se, either through
 4    the depletion of folate (Section 4.6.2) or the generation of reactive oxidant species (Section
 5    4.6.3).  Mouse livers actually have considerably higher hepatic tetrahydrofolate and total folate
 6    than rat or monkey liver. Harris et al. (2003) and Johlin et al. (1987) have shown that CAT
 7    activity in the embryo and VYS of rats and mice appear similar..
 8           Without positive identification of the actual moiety responsible for methanol-induced
 9    teratogenicity, MOA remains unclear. If the moiety is methanol, then it is possible that
10    generation of NADH during methanol oxidation creates an imbalance in other enzymatic
11    reactions.  Studies have shown that ethanol intake leads to a >100-fold increase in cellular
12    NADH, presumably due to ADHl-mediated reduction of the cofactor NAD+ to NADH
13    (Cronholm, 1987; Smith and Newman,  1959).  This is of potential importance because, for
14    example, ethanol intake has been shown to increase the in vivo and in vitro enzymatic reduction
15    of other endogenous compounds (e.g., serotonin) in humans (Svensson et al., 1999; Davis et al.,
16    1967).  In rodents, CAT-mediated methanol metabolism may obviate this effect; in humans,
17    however, methanol is primarily metabolized by ADH1.
18           If the teratogenic moiety of methanol is formaldehyde, then reactivity with protein
19    sulfhydryls and nonprotein sulfhydryls (e.g., GSH) or DNA protein cross-links may be involved.
20    Metabolic roles ascribed to ADH3, particularly regulation of S-nitrosothiol biology (Foster and
21    Stamler, 2004), could also be involved in the MOA.  Recently, Staab et al. (2008) have shown
22    that formaldehyde alters other ADH3-mediated reactions through cofactor recycling and that
23    formaldehyde alters levels of cellular S-nitrosothiol, which  plays a key role in cellular signaling
24    and many cellular functions and pathways (Hess et al.,  2005).
25           Studies such as those by Harris et al. (2004, 2003) and Dorman et al.  (1995, 1994)
26    suggest that formate is not the  metabolite responsible for methanol's teratogenic effects. The
27    former researchers suggest that formaldehyde is the proximate teratogen, and provide evidence
28    in support of that hypothesis. However, questions remain. Researchers in this  area have not yet
29    reported using a sufficient array of enzyme inhibitors to conclusively identify formaldehyde as
30    the proximate teratogen. Studies involving other inhibitors or toxicity studies carried out in
31    genetically engineered mice, while not devoid of confounders, might further inform regarding
32    the methanol MOA for developmental toxicity. Even if formaldehyde is ultimately identified as
33    the proximate teratogen, methanol would likely play a prominent role, at least in terms of
34    transport to the target tissue. The high reactivity of formaldehyde would limit its unbound and
35    unaltered transport as free formaldehyde from maternal to fetal blood (Thrasher and Kilburn,


                                                4-73        DRAFT-DO NOT CITE OR QUOTE

-------
 1    2001), and, as has been discussed, the capacity for the metabolism of methanol to formaldehyde
 2    is likely lower in the fetus and neonate versus adults (Section 3.3).

      4.6.2. Role of Folate Deficiency in the Developmental Toxicity of Methanol
 3          As discussed in Sections 3.1 and 4.1, humans and other primates are susceptible to the
 4    effects of methanol exposure associated with formate accumulation because they have lower
 5    levels of hepatic tetrahydrofolate-dependent enzymes that help in formate oxidation.
 6    Tetrahydofolate-dependent enzymes and critical pathways that depend on folate, such as purine
 7    and pyrimidine synthesis, may also play a role in the developmental toxicity of methanol.
 8    Studies of rats and mice fed folate-deficient diets have identified adverse effects on reproductive
 9    performance, implantation, fetal growth and developmental defects, and the inhibition of folate
10    cellular transport has been associated with several  developmental abnormalities, ranging from
11    neural tube defects to neurocristopathies such as cleft-lip and cleft-palate, cardiacseptal defects,
12    and eye defects (Antony, 2007). Folate deficiency has been shown to exacerbate some aspects of
13    the developmental toxicity of methanol in mice (see discussion of Fu et al., 1996, and Sakanashi
14    et al., 1996, in Section 4.3.1) and rats (see discussion of Aziz et al., 2002,  in Section 4.4.1).
15          The studies in mice focused on the influence of FAD on the reproductive and skeletal
16    malformation effects of methanol.  Sakanashi  et al. (1996) showed that dams exposed to
17    5 g/kg-day methanol on GD6-GD15 experienced a threefold increase in the percentage of litters
18    affected by cleft palate and a 10-fold increase in the percentage of litters affected by exencephaly
19    when fed a FAD (resulting in a 50% decrease in liver folate) versus a FAS diet. They speculated
20    that the increased methanol effect from FAD diet could have been due to an increase in tissue
21    formate or a critical reduction in conceptus folate concentration immediately following the
22    methanol exposure.  The latter appears more likely, given the high levels of formate needed to
23    cause embryotoxicity (Section 4.3.3) and the decrease in conceptus folate that is observed within
24    2 hours of GD8 methanol exposure (Dorman et al., 1995). Fu et  al. (1996) confirmed the
25    findings of Sakanashi et al. (1996) and also  determined that the maternal FAD diet had a much
26    greater impact on fetal liver folate than maternal liver folate levels.
27          The rat study of Aziz et al. (2002) focused on the influence of FAD on  the developmental
28    neurotoxicity of methanol.  Experiments by  Aziz et al. (2002) involving Wistar rat dams and
29    pups  exposed to methanol during lactation provide evidence that methanol exposure during this
30    postnatal period affects the developing brain.  These effects (increased spontaneous locomotor
31    activity, decreased conditioned avoidance response, disturbances in dopaminergic and
32    cholinergic receptors and increased expression of GAP-43 in the  hippocampal region) were more
33    pronounced in FAD as compared to FAS rats.  This suggests that folic acid may play a role in
34    methanol-induced neurotoxicity. These results do  not implicate any particular proximate
                                                4-74       DRAFT-DO NOT CITE OR QUOTE

-------
 1    teratogen, as folate deficiency can increase levels of both methanol, formaldehyde and formate
 2    (Medinsky et al., 1997). Further, folic acid is used in a number of critical pathways such as
 3    purine and pyrimidine synthesis. Thus, alterations in available folic acid, particularly to the
 4    conceptus, could have significant impacts on the developing fetus apart from the influence it is
 5    presumed to have on formate removal.

      4.6.3. Methanol-Induced Formation of Free Radicals, Lipid Peroxidation, and Protein
      Modifications
 6          Oxidative stress in mother and offspring has been suggested to be part of the teratogenic
 7    mechanism of a related alcohol, ethanol. Certain reproductive and developmental effects (e.g.,
 8    resorptions and malformation rates) observed in Sprague-Dawley rats following ethanol
 9    exposure were reported to be ameliorated by antioxidant (Vitamin E) treatment (Wentzel et al.,
10    2006; Wentzel and Eriksson, 2006). A number of studies have examined markers of oxidative
11    stress associated with methanol exposure.
12          Skrzydlewska et al. (2005) provided inferential evidence for the effects of methanol on
13    free radical formation, lipid peroxidation, and protein modifications, by studying the protective
14    effects of N-acetyl cysteine and the Vitamin E derivative, U83836E, in the liver of male Wistar
15    rats exposed to the compound via gavage. Forty-two rats/group received a single oral gavage
16    dose of either saline or 50% methanol.  This provided a dose of approximately 6,000 mg/kg,  as
17    calculated by the authors.  Other groups of rats received the same concentration of methanol, but
18    were also injected intraperitoneally with either N-acetylcysteine or U-83836E.  N-acetylcysteine
19    and U-83836E controls were also included in the study design. Animals in each group were
20    sacrificed after 6, 14,  and 24 hours or after 2, 5, or 7 days. Livers were rapidly  excised for
21    electron spin resonance (ESR) analysis, and 10,000 x g supernatants were used to measure GSH,
22    malondialdehyde, a range of protein parameters, including free amino and sulfhydryl groups,
23    protein carbonyls, tryptophan, tyrosine, and bityrosine, and the activity  of cathepsin B.
24          Skrzydlewska et al. (2005) provided data that showed an increase in an ESR signal at
25    g = 2.003 in livers harvested 6 and 12 hours after methanol exposure. The signal, thought to be
26    indicative of free radical formation, was opposed by N-acetylcysteine and U83836E.  Other
27    compound-related changes included: 1) a significant decrease in GSH levels that was most
28    evident in rats sacrificed 12 and 24 hours after exposure; 2) increased concentrations in the lipid
29    peroxidation product, malondialdehyde (by a maximum of 44% in the livers of animals
30    sacrificed 2 days after exposure); 3) increased specific concentrations of protein carbonyl groups
31    and bityrosine; but 4) reductions in the specific level of tryptophan. Given the ability of N-
32    acetylcysteine and U83836E to oppose these changes, at least in part, the authors  speculated that
33    a number of potentially harmful changes may have occurred as a result  of methanol exposure.

                                               4-75        DRAFT-DO NOT CITE OR QUOTE

-------
 1    These include free radical formation, lipid peroxidation, and disturbances in protein structure.
 2    However, it is unclear whether or not the metabolites of methanol, formaldehyde, and/or formate,
 3    were involved in any of these changes.
 4          Rajamani et al. (2006) examined several oxidative stress parameters in male Wistar rats
 5    following methotrexate-induced folate deficiency.  Compared to controls, the levels of free
 6    radical scavengers SOD, CAT, GSH peroxidase, oxidized GSH, protein carbonyls, and lipid
 7    peroxidation were elevated in several regions of the brain, with greater increases observed in the
 8    MTX-methanol-treated animals than in the MTX-alone group. The level of GSH and protein
 9    thiols was decreased in all regions of the brain, with a greater decrease observed in the MTX-
10    methanol-treated animals than MTX-treated animals.
11          Dudka (2006) measured the total antioxidant status (TAS) in the brain of male Wistar rats
12    exposed to a single oral gavage dose of methanol at 3 g/kg.  The animals were kept in a nitrous
13    oxide atmosphere (N2O/O2) throughout the experiment to reduce intrinsic folate levels, and
14    various levels of ethanol and/or fomepizole (as ADH antidotes) were administered i.p. after
15    4 hours. Animals were sacrificed after 16 hours, the brains homogenized, and the TAS
16    determined spectrophotometrically.  As illustrated graphically by the author, methanol
17    administration reduced TAS in brain irrespective of the presence of ADH antidotes.  The author
18    speculated that, while most methanol is metabolized in the liver, some may also reach the brain.
19    Metabolism to formate might then alter the NADH/NAD+ ratio resulting in an increase in
20    xanthine oxidase activity and the formation of the superoxide anion.
21          Parthasarathy et al. (2006b) investigated the extent of methanol-induced oxidative stress
22    in rat lymphoid organs. Six male Wistar rats/group received 2,370 mg/kg methanol (mixed 1:1
23    with saline) injected i.p. for 1,  15 or 30 days. A control group received a daily i.p. injection of
24    saline for 30 days. At term, lymphoid organs such as the spleen, thymus, lymph  nodes, and bone
25    marrow were excised, perfused with saline, then homogenized to obtain supernatants in which
26    such indices of lipid peroxidation as malondialdehyde, and the activities of CAT, SOD, and GSH
27    peroxidase were measured. Parthasarathy et al. (2006b) also measured the concentrations of
28    GSH and ascorbic acid (nonenzymatic antioxidants) and the serum concentrations of a number of
29    indicators of liver and kidney function,  such as ALT, AST, blood urea nitrogen (BUN), and
30    creatinine.
31          Table 4-23 shows the time-dependent changes in serum liver and kidney function
32    indicators that resulted from methanol administration. Treatment with methanol  for increasing
33    durations resulted in increased serum ALT and AST activities and the concentrations of BUN and
34    creatinine.
                                               4-76        DRAFT-DO NOT CITE OR QUOTE

-------
         Table 4-23. Time-dependent effects of methanol administration on serum liver and
         kidney function, serum ALT, AST, BUN, and creatinine in control and experimental
         groups of male Wistar rats
Parameters
ALT (umoles/min-mg)
AST (umoles/min-mg)
BUN (mg/L)
Creatinine (mg/L)
Methanol administration (2,370 mg/kg)
Control
29.0 ±2.5
5.8 ±0.4
301 ±36
4.6 ±0.3
Single dose
31.4±3.3
6.4 ±0.3
332 ±29
4.8 ±0.3
15 days
53.1±2.3a
9.0±1.2a
436±35a
5.6±0.2a
30 days
60.4±2.8a
13.7±1.2a
513±32a
7.0±0.4a
    Values are means ± S.D. of 6 animals.
    a/> < 0.05 versus controls.
    Source: Parthasarathy et al. (2006b) (adapted).
         Table 4-24. Effect of methanol administration on male Wistar rats on malondialdehyde
         concentration in the lymphoid organs of experimental and control groups and the
         effect of methanol on antioxidants in spleen
Parameters
Methanol administration (2,370 mg/kg)
Control
Single dose
15 days
30 days
Malondialdehyde in lymphoid organs
Spleen
Thymus
Lymph nodes
Bone marrow
2.62 ±0.19
3.58 ±0.35
3. 15 ±0.25
3. 14 ±0.33
4.14±0.25a
5.76±0.36a
5.08±0.24a
4.47±0.18a
7.22±0.31a
9.23±0.57a
8.77±0.57a
7.20 ± 0.42a
9.72±0.52a
11.6±0.33a
9.17±0.67a
9.75±0.56a
Antioxidant levels in spleen
SOD (units/mg protein)
CAT (umoles H2O2
consumed/min-mg protein
GPx (ug GSH
consumed/min-mg protein)
GSH (ug/mg protein)
Vit C (ug/mg protein)
2.40 ±0.16
35.8 ±2.77
11.2 ±0.60
2.11±0.11
0.45 ±0.04
4.06±0.19a
52.5±3.86a
20.0±1.0a
3.75±0.15a
0.73±0.05a
1.76±0.09a
19.1±1.55a
7.07±0.83a
1.66±0.09a
0.34±0.18a
1.00±0.07a
10.8±1.10a
5.18±0.45a
0.89±0.04a
0.11±0.03a
    Values are means ± S.D. of six animals.
    a/> < 0.05, versus controls.
    Source: Parthasarathy et al. (2006b) (adapted).

1          Table 4-24 gives the concentration of malondialdehyde in the lymphoid organs of control
2   and experimental groups, and, as an example of all tissue sites examined, the levels of enzymatic
3   and nonenzymatic antioxidants in spleen.  The results show that malondialdehyde concentrations
4   were time-dependently increased at each tissue site and that, in spleen as an example of all the
5   lymphoid tissues examined, increasing methanol administration resulted in lower levels of all
6   antioxidants examined compared to controls. Parthasarathy et al. (2006b) concluded that
                                              4-77
DRAFT-DO NOT CITE OR QUOTE

-------
 1    exposure to methanol may cause oxidative stress by altering the oxidant/antioxidant balance in
 2    lymphoid organs in the rat.

      4.6.4. Exogenous Formate Dehydrogenase as a Means of Detoxifying the Formic Acid that
      Results from Methanol Exposure
 3          In companion reports, Muthuvel et al. (2006a, 2006b) used 6 male Wistar rats/group to
 4    test the ability of exogenously-administered formate dehydrogenase (FD) to reduce the serum
 5    levels of formate that were formed when 3 g/kg methanol was administered i.p. to rats in saline.
 6    In the first experiment, purified FD (from Candida boitinif) was administered by i.v. conjugated
 7    to the N-hydroxysuccinimidyl ester of monomethoxy polyethylene glycol propionic acid
 8    (PEG-FD) (Muthuvel et al., 2006a).  In the second, rats were administered FD-loaded
 9    erythrocytes (Muthuvel et al., 2006b). In the former case, some groups of rats were made folate
10    deficient by means of a folate-depleted diet; in the latter, folate deficiency was brought about by
11    i.p. administration of methotrexate. In some groups, the rats received an infusion of an
12    equimolar mixture of carbonate and bicarbonate (each at 0.33 mol/L) to correct the formate-
13    induced  acidosis. As illustrated by the authors, methanol-exposed rats receiving a folate-
14    deficient diet showed significantly higher levels of serum formate than those receiving a folate-
15    sufficient diet.  However, administration of native or PEG-FD reduced serum formate in
16    methanol-receiving folate-deficient rats to levels seen in animals receiving methanol and the
17    folate-sufficient diet.
18          In the second report, Muthuvel et al. (2006b) carried out some preliminary experiments to
19    show that hematological parameters of normal, reconstituted but unloaded, and reconstituted and
20    FD-loaded erythrocytes, were similar. In addition, they showed that formate levels of serum
21    were reduced in vitro in the presence of FD-loaded erythrocytes.  Expressing blood formate
22    concentration in mmol/L at the 1-hour time point after carbonate/bicarbonate and enzyme-loaded
23    erythrocyte infusion via the tail vein, the concentration was reduced from 10.63 ±1.3 (mean ±
24    S.D.) in methanol and methotrexate-receiving controls to 5.83 ± 0.97 (n = 6).  This difference
25    was statistically significant at the/? < 0.05 level. However, FD-loaded erythrocytes were less
26    efficient at removing formate in the absence of carbonate/bicarbonate. Effective elimination of
27    formate appears to require an optimum pH for the FD activity in the enzyme-loaded
28    erythrocytes.

      4.6.5. Mechanistic Data Related to the Potential Carcinogenicity of Methanol
      4.6.5.1. Gen otoxicity
29          The genotoxicity/mutagenicity of methanol has not been extensively studied, but the
30    results of those studies that have thus far have been mostly negative. For example, in a survey of
                                               4-78        DRAFT-DO NOT CITE OR QUOTE

-------
 1   the capacity of 71 drinking water contaminants to induce gene reversion in the Ames test,
 2   Simmon et al. (1977) listed methanol as one of 45 chemicals that gave negative results with
 3   Salmonella typhimurium strains TA98, 100, 1535, 1537, and 1538, irrespective of the presence
 4   or absence of metabolic activation (an S9 microsomal fraction). This result was confirmed by
 5   DeFlora et al. (1984) and in NEDO (1987) for the same strains of Salmonella. DeFlora et al.
 6   (1984) also found methanol to be negative for induction of DNA repair in E. coli strains WP2,
 7   WP2 (uvrA, polA~), and CM871 (uvrA, recA, lexA~), again irrespectively of the presence or
 8   absence of S9.
 9          Abbondandolo et al. (1980) used a ade6-60/radlO-198,h~ strain of Schizosaccharomyces
10   pombe (PI strain) to determine the capacity of methanol and other solvents to induce forward
11   mutations. Negative results were obtained for methanol, irrespective of metabolic activation
12   status. In other genotoxicity/mutagenicity studies of methanol using fungi, Griffiths (1981)
13   reported methanol to be negative for the induction of aneuploidy in Neurospora crassa. By
14   contrast, weakly positive results for the compound were obtained by Crebelli et al. (1989) for the
15   induction of chromosomal malsegregation in the diploid strain PI of Aspergillus nidulans.
16          In an  extensive review of the capacity of a wide range of compounds to induce
17   transformation in mammalian cell lines, Heidelberger et al. (1983) reported methanol to be
18   negative in Syrian hamster embryo (SHE) cells. It also did not enhance the transformation of
19   SHE cells by Simian adenovirus. However, McGregor et al. (1985) reported in an abstract that a
20   statistically significant increase in forward mutations in the mouse lymphoma L5178Y tk+/tk" cell
21   line occurred at a concentration of 7.9 mg/mL methanol in the presence of S9.
22          The capacity of methanol to bring about genetic changes in human cell lines was
23   examined by Ohno et al. (2005), who developed a system in which the chemical activation of the
24   p53R2 gene was assessed by the incorporation of a/>5.iR2-dependent luciferase reporter gene
25   into two human cell lines, MCF-7 and HepG2. Methanol, among 80 chemicals tested in this
26   system, gave negative results. NEDO (1987) used Chinese hamster lung (CHL) cells to monitor
27   methanol's capacity to induce 1) forward mutations to azaguanine, 6-thioguanine, and ouabain
28   resistance, and 2) chromosomal aberrations (CA) though with negative results throughout.
29   However, methanol did display some capacity to induce sister chromatid exchanges (SCE) in
30   CHL cells, since the incidence of these lesions at the highest concentration (28.5 mg/mL) was
31   significantly  greater than in controls (9.41 ± 0.416 versus 6.42 ± 0.227 [mean ± SE per  100
32   cells]).
33          In an  in vivo experiment examining the genotoxicity/mutagenicity of methanol,  Campbell
34   et al. (1991) exposed 10 male C57BL/6J mice/group to 0, 800, or 4,000 ppm (0, 1,048, and
35   5,242 mg/m3) methanol, 6 hours/day, for 5 days. At sacrifice, blood cells were examined for the
36   formation of micronuclei (MN). Excised lung cells for SCE, CA and MN, and excised testicular

                                               4-79         DRAFT-DO NOT CITE OR QUOTE

-------
 1   germ cells were examined for evidence of synaptonemal damage, in each case with negative
 2   results.
 3          There was no evidence of methanol-induced formation of MN in the blood of fetuses or
 4   pregnant CD-mice when the latter were gavaged twice daily with 2,500 mg/kg methanol on
 5   GD6-GD10 (Fu et al., 1996).  The presence of marginal or adequate amounts of folic acid in the
 6   diet of the dams did not affect MN formation. NEDO (1987) carried out an in vivo MN test in
 7   6 male SPF mice/group who received a single gavage dose of 1,050, 2,110, 4,210, and 8,410
 8   mg/kg methanol. Twenty-four hours later, 1,000 cells were counted for MN in bone marrow
 9   smears. No compound-related effects on MN incidence were observed.  Table 4-25 provides a
10   summary of the genotoxicity/mutagenicity studies of methanol.

          Table 4-25. Summary of genotoxicity studies of methanol
Test system
Cell/strain
Result
Reference
Comments
In vitro tests
Gene reversion/
S. typhimurium
DNA repair/It, coli
Forward mutations/
S. pombe
Aneuploidy/
N. crassa
Chromosomal
malsegregation/
A. nidulans
Forward
mutations/Mouse
lymphoma cells
Forward
mutations/Chinese
hamster lung cells
Chromosomal
aberrations/Chinese
hamster lung cells
Sister chromatid
exchanges/Chinese
hamster lung cells
TA98; TA100; TA1535,
TA1537, TA1538
TA98; TA100; TA1535,
TA1537, TA1538
TA98; TA100; TA1535,
TA1537, TA1538
WP2, WP2 (uvrA,
polA'),CM&ll(uvrA', recA~,
lexA)
PI
(ade6-60/radl 0-198, K)
(arg-1, ad-3A, ad-3B, nic-
2, to/, C/c, D/d, E/e)
PI (diploid)
L5178Ytk+/tk~
to azaguanine, 6-
thioguanine and ouabain
resistance


- (+S9); - (-S9)
- (+S9); - (-S9)
- (+S9); - (-S9)
- (+S9), - (-S9)
-(+S10), -(-S10)
- (S9 status not
reported)
+ (S9 status not
reported)
+ (+S9), ND (-S9)
- (-S9), ND (+S9)
- (-S9), ND (+S9)
+ (-S9), ND (+S9)
Simmon etal. (1977)
De Flora etal. (1984)
NEDO (1987)
DeFlora etal. (1984)
Abbondandolo et al.
(1980)
Griffiths (1981)
Crebelli etal. (1989)
McGregor et al.
(1985)
NEDO (1987)
NEDO (1987)
NEDO (1987)




Molecular activation
used a 10,000 xg(S10)
supernatant from liver
of induced Swiss mice


Results reported in an
abstract



                                              4-80
DRAFT-DO NOT CITE OR QUOTE

-------
Test system
Genetic activation/
human cell lines
Cell transformation/
Syrian hamster
embryo cells
Cell/strain
MCF-7 and HepG2
containing ap53R2-
dependent luciferase
reporter gene
with/without
transformation by Simian
adenovirus
Result
- (-S9), ND (+S9)
- (-S9), ND (+S9)
- (-S9), ND (+S9)
Reference
Ohno et al. (2005)
Heidelberger et al.
(1983)
Comments

Review
In vivo tests
Mouse/MN
formation
Mouse/SCEs
Mice/CA
Mouse/synaptonem
al damage
Mouse/MN
formation
C57BL/6J
(Blood cells)
C57BL/6J
(Lung cells)
C57BL/6J
(Lung cells)
C57BL/6J
(Lung cells)
C57BL/6J
(Testicular germ cells)
CD-I
(Blood cells)
SPF
(Bone marrow cells)
-
-
-
-
-
-
-
Campbell et al.
(1991)
Campbell et al.
(1991)
Campbell et al.
(1991)
Campbell et al.
(1991)
Campbell et al.
(1991)
Fuetal. (1996)
NEDO (1987)
Molecular activation
not applicable
Molecular activation
not applicable
Molecular activation
not applicable
Molecular activation
not applicable
Molecular activation
not applicable
Molecular activation
not applicable
Molecular activation
not applicable
     ND = not determined.

     4.6.5.2. Lymphoma Responses Reported in ERF Life span Bioassays of Compounds
     Related toMethanol, Including an Analogue (Ethanol), Precursors (Aspartame and Methyl
     Tertiary Butyl Ether), and a Metabolite (Formaldehyde)
 1          The ERF or the European Foundation of Oncology and Environmental Sciences have
 2   conducted nearly 400 experimental bioassays on over 200 compounds/agents, using some
 3   148,000 animals over nearly 4 decades.  Of the over 200 compounds tested by ERF,52 8 have
 4   been associated with an increased incidence of hemolymphoreticular tumors in Sprague-Dawley
 5   rats, suggesting that it may be a rare and potentially species/strain-specific finding. These eight
 6   chemicals are: methanol, formaldehyde, aspartame, MTBE, DIPE, TAME, mancozeb, and
 7   toluene. Methanol, formaldehyde, aspartame, and MTBE share a common metabolite,
 8   formaldehyde, and DIPE, TAME, methanol and MTBE are all gasoline-oxygenate additives
 9   (Caldwell  et al., 2008).
10          With the exception of a positive study for malignant lymphomas in Swiss Webster mice
11   exposed to methanol (Apaja, 1989), lymphoma responses have not been reported by other
12   institutions performing long-term testing of these chemicals in various strains of rats, including
     52 While ERF has tested over 200 chemicals in 398 long-term ERF bioassays, only 112 of their bioassays have been
     published to date (Caldwell, et al., 2008). The extent to which the unpublished studies are documented varies.
                                              4-81       DRAFT-DO NOT CITE OR QUOTE

-------
 1    formaldehyde inhalation studies in F344 (Kerns et al., 1983; Kamata et al., 1987)53 and Sprague-
 2    Dawley (Albert et al., 1982; Sellakumar et al., 1985) rats, formaldehyde oral studies in Wistar
 3    rats (Til et al., 1989; Tobe et al., 1989), toluene oral  studies in F344 rats (NTP, 1990), MTBE
 4    inhalation studies in F344 rats (Chun et al., 1992), aspartame oral studies in Wistar (Ishii et al.,
 5    1981) and Sprague-Dawley (Molinary, 1984) rats, and methanol inhalation studies in F344 rats
 6    (NEDO, 1987, 1985/2008b).  Several differences in study design may contribute to the
 7    differences in responses observed across institutions, particularly study duration and test animal
 8    strain.  Fischer-344 rats have a high background of mononuclear cell leukemia (20% in control
 9    females)54 and a very low background rate of "lymphoma" (0% in control females) at 104 weeks
10    (NTP, 2006). In contrast, Sprague-Dawley rats from NTP studies exhibit a low background rate
11    of "leukemias" (0.8% in control females) and a higher background rate of "lymphomas" (1.08%
12    in control females) at  104 weeks (NTP, 2006). Similarly, Chandra et al. (1992) report a
13    background level  of 1.6% for malignant lymphocytic lymphomas in female control Sprague-
14    Dawley rats for 17 2-year carcinogenicity studies.
15           In lifetime studies of Sprague-Dawley rats at ERF, the overall incidence of
16    lymphomas/leukemias has been reported to be 13.3% (range, 4.0-25.0%) in female historical
17    controls (2,274 rats) and 20.6% (range, 8.0-30.9%)  in male historical controls (2,265 rats)
18    (Soffritti et al., 2007).  The difference in background rates reported by ERF versus other labs for
19    this tumor type could be due to  differences in study  duration, differences in tumor classification
20    systems, and/or misdiagnoses due to confounding effects (see discussion in Section 4.2.1.3). A
21    high background incidence can  increase the difficulty of detecting chemically related responses
22    (Melnick et al., 2007), and the background rate reported by ERF for this tumor type is considered
23    to be high relative to other tumor types and relative to the background rate for this tumor type in
24    Sprague-Dawley rats from other laboratories (EFSA, 2006; Cruzan,  2009).55 However, it is in a
25    range that can be  considered reasonable for studies that employ a large  number of animals
26    (Leakey et al., 2003; Caldwell et al., 2009).
27           Thus, with respect to the identification of hemolymphoreticular carcinogenic responses,
28    life span studies of Sprague-Dawley rats performed  by ERF may be  more sensitive than the
29    2-year studies of Fischer 344 (F344) strain of rats used by NTP (1990) and NEDO (1987,
      53 Though Kerns et al. (1983) did not report a positive response for lymphoma, a survival-adjusted analysis of the
      data from this study indicates a statistically significant trend in female rat mononuclear cell leukemia (p = 0.0056)
      and a nearly significant increase in female mouse lymphoma (p = 0.06). In the Kamata et al., (1987) study, only a
      small percentage of the original 32 rats/group survived to the end of the study (28 months) due largely to interim
      sacrifices (5/group) at 12, 18 and 24 months.
      54 Due to this and other health concerns, NTP transitioned to the use of Wistar rats in 2008, and more recently has
      adopted Sprague-Dawley rats as the rat model for NTP studies due to the reproductive capability and size of Wistar
      rats (http://ntp.niehs.nih.gov/go/29502).
      55 Cruzan (2009) reports that the incidences of total cancers derived from bloodforming cells, designated as
      hemolymphoreticular tumors by Ramazzini pathologists, is consistently about four times higher than the incidences
      of such tumors in SD rats recorded in the Charles River Laboratory historical database (CRL database).
                                                 4-82        DRAFT-DO NOT CITE OR QUOTE

-------
 1    1985/2008b). The results of ERF studies of the carcinogenic potential of methanol, MTBE, and
 2    formaldehyde and related chemicals, ethanol and aspartame, are summarized in this section.
 3    4.6.5.2.1. Ethanol.  In a study that was reported in the same article that described the
 4    carcinogenic responses of Sprague-Dawley rats to methanol, Soffritti et al. (2002a) exposed
 5    110 Sprague-Dawley rats/sex/group to ethanol in drinking water at concentrations of 0 or 10%
 6    (v/v) beginning at 39 weeks of age and ending at natural death, and including a single breeding
 7    cycle.  Various numbers of the offspring (30 male controls, 39 female controls, 49 exposed
 8    males, and 55  exposed females) were exposed to ethanol in drinking water at the same
 9    concentrations as their parents.  The experiment concluded with the death of the last offspring at
10    179 weeks of age. Animals were examined for the same toxicological parameters as those
11    described for methanol, and organs and tissues were grossly and histopathologically examined at
12    necropsy. Soffritti et al. (2002a) reported that food and drinking water intake were lower in
13    exposed animals compared  to controls but that body weight changes were similar among the
14    groups.56 There were no compound-related clinical signs of toxicity and no differences in
15    survival  rates among the groups. While there were apparently no nononcogenic pathological
16    changes  evident on gross inspection or histopathologic examination, a number of benign and
17    malignant tumors were considered by the authors to be compound-related. Compared to
18    controls, these included increased incidences of: 1) total malignant tumors in male and female
19    breeders (145/220 versus 99/220)  and offspring (49/69 versus 54/104); 2) total malignant tumors
20    per 100 animals in female breeders (130 versus 60.9) and offspring (164.1 versus 96.4); 3)
21    carcinomas of the head and neck, especially to the oral cavity, lips and tongue in  male and
22    female breeders (27/220 versus 5/220) and offspring (26/69 versus 5/104); 4) squamous cell
23    carcinomas of the forestomach in male and female breeders (5/220 versus 0/220) and offspring
24    (2/69 versus 0/104); 5) interstitial  cell adenomas of the testis in male breeders (23/110 versus
25    9/110) and offspring (4/30 versus 4/49); 6) Sertoli-Leydig cell tumors in ovaries of female
26    offspring (3/39 versus 1/55); 7) adenocarcinomas of the uterus in female breeders (9/110 versus
27    2/110) and offspring (8/39 versus 6/55); 8) pheochromoblastomas in male and female breeders
28    (13/220 versus 4/220) and offspring (4/69 versus 2/104); and 9) osteosarcomas in male and
29    female breeders ([for the head] 14/220 versus 4/220) and offspring ([for the head] 10/69 versus
30    7/104). Notably,  Soffritti et al. (2002a) did not observe increases in any of the lymphoma
31    responses reported in their methanol bioassay. Incidence data for these responses and their
      statistical significance compared to controls are shown in Table 4-26.
      56 Test animals were likely receiving calories from ethanol exposure.
                                               4-83        DRAFT-DO NOT CITE OR QUOTE

-------
          Table 4-26. Incidence of carcinogenic responses in Sprague-Dawley rats exposed to
          ethanol in drinking water for up to 2 years
Tissues/affected sites
Total malignant tumors
Oral cavity (carcinomas)
Forestomach (squamous
cell carcinomas)
Testis (interstitial cell
adenomas)
Sertoli-Leydig cell tumors
(ovary)
Uterus (adenocarcinomas)
Head (osteosarcomas)
Adrenal gland
(pheochromoblastomas)
Total malignant tumors per
100 animals
Concentration in percent (v/v)
Male (breeders)
0
51/110
3/110
0/110
9/110


0/110
3/110
61.8
10
66/110
15/1 10b
2/110
23/1 10d


8/110
9/110
89.1b
Female (breeders)
0
48/110
2/110
0/110

1/110
2/110
4/110
1/110
60.9
10
79/1 10b
12/110
3/110

2/110
9/1 10C
6/110
4/110
130b
Male (offspring)
0
23/49
2/49
0/49
4/49


4/49
1/49
61.2
10
23/303
10/30b
1/30
4/30


6/30
4/30
136.7b
Female (offspring)
0
31/55
3/55
0/55

0/55
6/55
3/55
1/55
96.4
10
26/39
16/39b
1/39

3/39
8/39
4/39
0/39
164.1b
     a/> < 0.05 using the ^ test, as calculated by the authors.
     bp < 0.01 using the %2 test, as calculated by the authors.
     °p < 0.05 using Fisher's exact test, as calculated by the reviewers.
     dp < 0.01 using Fisher's exact test, as calculated by the reviewers.
     Source: Soffritti et al. (2002a).
 1   4.6.5.2.2. Aspartame. Soffritti et al. (2006, 2005) reported the results of a cancer bioassay on
 2   the artificial sweetener aspartame. The study has potential relevance to the carcinogenicity of
 3   methanol because aspartame has been shown to be metabolized to aspartic acid, phenylalanine,
 4   and methanol in the GI tract prior to absorption into systemic circulation. In the study, aspartame
 5   (>98% purity) was given to 100 or 150 Sprague-Dawley rats/sex/group in feed at dietary
 6   concentrations of 0, 80, 400, 2,000, 10,000, 50,000, and 100,000 ppm.  The  authors reported
 7   these concentrations to be equivalent to approximate daily doses of 0, 4, 20, 100,  500, 2,500, and
 8   5,000 mg/kg-day, respectively, under the conditions of the study. Animals were maintained until
 9   their "natural death," with the in-life phase of the experiment concluding with the death of the
10   last animal at 151 weeks.  All animals were monitored for body weight, food and water
11   consumption. At death, animals were examined grossly and given a complete histopathological
12   examination.
13          Soffritti et al. (2006, 2005) reported that there were no differences among the groups in
14   mean body weight, survival, or daily water consumption. However, there appeared to be a dose-
15   related reduction in food consumption in both male and female rats.
                                                4-84        DRAFT-DO NOT CITE OR QUOTE

-------
 1          The principal histopathological finding was an increased incidence of lymphomas and
 2    leukemias in female rats, a response reported by the authors to be statistically significant
 3    compared to concurrently exposed controls (Table 4-27) and greater than the range of overall
 4    incidence of lymphomas and leukemias in historical controls at the ERF (13.4% [range, 7.0-
 5    18.4] in females and 21.8% [range, 8.0-30.9] in males). Among the hemolymphoreticular
 6    neoplasms observed, the most frequent type observed was lympho-immunoblastic lymphoma.
 7    The authors concluded that aspartame causes a "dose-related statistically significant increase in
 8    lymphomas and leukemias in females at dose levels very near those to which humans can be
 9    exposed." They postulated that an increase in the incidence of lymphomas and leukemias could
10    be associated with the formation of either methanol or formaldehyde.  Other potentially
11    compound-related effects of aspartame were (1) an increase in combined dysplastic hyperplasias,
12    papillomas, and carcinomas of the renal  pelvis and ureter, and (2) an increasing trend in the
13    formation of malignant schwannomas in peripheral nerves (Table 4-28).
14          The European Commission asked the European Food Safety Authority (EFSA) to assess
15    the study and review all ERF findings related to aspartame.  An EFSA review panel assessed the
16    study and considered additional unpublished data provided to it by the ERF.  In their report,
17    EFSA (2006) concluded that the Soffritti et al. (2005, 2006) study had flaws that brought into
18    question the reported findings. The review panel noted the high background of chronic
19    inflammatory changes in the lung and other vital organs.  These background inflammatory
20    changes were thought to contribute significant uncertainty to the interpretations  of the study. In
21    fact, the review panel concluded that most of the documented changes, in particular, the apparent
22    compound-related increase in lymphomas and leukemias, may have been incidental findings and,
23    therefore, unrelated to aspartame.

          Table 4-27. Incidence of lymphomas and leukemias in Sprague-Dawley rats exposed to
          aspartame via the diet
Group
I
II
III
IV
V
VI
VII
ppm in feed
0
80
400
2,000
10,000
50,000
100,000
Dose (mg/kg-day)
0
4
20
100
500
2,500
5,000
Lymphomas/leukemias (incidence and %)
Male
31/150(21)
23/150 (15)
25/150 (17)
33/150 (22)
15/100 (15)
20/100 (20)
29/100 (29)
Female
13/150 (9)
22/150 (15)
30/150b (20)
28/150a(19)
19/100a(19)
25/100b (25)
25/100b (25)
      *p < 0.05 using the poly-k test.
      V < 0.01 using the poly-k test.
      Source: Soffritti et al. (2006, 2005).
                                               4-85
DRAFT-DO NOT CITE OR QUOTE

-------
          Table 4-28. Incidence of combined dysplastic hyperplasias, papillomas and carcinomas
          of the pelvis and ureter and of malignant schwannomas in peripheral nerve in
          Sprague-Dawley rats exposed to aspartame via the diet
Group
I
II
III
IV
V
VI
VII
ppm in feed
0
80
400
2,000
10,000
50,000
100,000
Dose
(mg/kg-day)
0
4
20
100
500
2,500
5,000
Incidence and %
Combined hyperplasias,
papillomas and carcinomas
of the pelvis and ureter
Male
1/150 (0.7)
3/149 (2)
5/149 (3)
5/150 (3)
3/100 (3)
3/100 (3)
4/100 (4)
Female
2/150c (1.3)
6/150 (4)
9/1 50a (6)
10/1503 (7)
10/100b (10)
10/99b (10)
15/100b (15)
Peripheral nerve malignant
schwannomas
Male
1/150C (0.7)
1/150 (0.7)
3/150 (2)
2/150(1.3)
2/100 (2)
3/100 (3)
4/100 (4)
Female
0/150 (0)
2/150(1.3)
0/150 (0)
3/150 (2)
1/100 (1)
1/100 (1)
2/100 (2)
      ap < 0.05 using the poly-k test.
      bp < 0.01 using the poly-k test.
      °p < 0.05 using the Cochran-Armitage test for trend.
      Source: Soffritti et al. (2006).

 1          In their conclusions, the EPS A review panel took note of negative results of 2-year
 2    carcinogenic studies of aspartame (Ishii et et al., 1981; Ishii, 1981; NTP, 2003) and of the
 3    findings of a recent epidemiological study carried out by the US National Cancer Institute (NCI,
 4    2006).
 5          In an effort to further clarify these issues, Soffritti et al. (2007) reported the results of
 6    another lifetime study of aspartame in which 95 controls and 70 Sprague-Dawley rats/sex/group
 7    were exposed, first in utero, then via the diet, to aspartame at concentrations of 0, 400, or
 8    2,000 ppm (mg/kg) of feed. The authors assumed an average food consumption of 20 g/day and
 9    an average body weight (males and females) of 400 g, thereby deriving average target doses of 0,
10    40, and 200 mg/kg-day.  Soffritti et al. (2007) began administering the aspartame-supplemented
11    feed to the dams on GDI2; and offspring received feed containing aspartame at the appropriate
12    concentration from weaning until natural death. Animals were observed three times daily
13    Monday-Friday, and twice daily on Saturdays, Sundays, and holidays.  This regimen was both to
14    monitor clinical signs and to reduce the possibility of decedents undergoing autolysis before
15    discovery. As described by the authors, all deceased animals were refrigerated, then necropsied
16    no more than 19 hours after discovery.
17          Food and drinking water consumption was monitored once/day. Beginning at 6 weeks of
18    age, individual body weights were recorded once a week for 13 weeks, then every 2 weeks until
19    natural death. All animals were examined grossly every 2 weeks.  After necropsy, tissues and
20    organs were sampled for histopathological processing and microscopic examination (including
                                               4-86         DRAFT-DO NOT CITE OR QUOTE

-------
 1    skin and subcutaneous tissue, mammary gland, brain, pituitary, Zymbal's gland, salivary gland,
 2    Harderian gland, cranium, tongue, thyroid, parathyroid, pharynx, larynx, thymus and mediastinal
 3    lymph nodes, trachea, lung and main stem bronchi, heart, diaphragm, liver, spleen, pancreas,
 4    kidney, adrenal gland, esophagus, stomach, intestine, urinary bladder, prostate, vagina, gonads,
 5    interscapular brown fat pads, subcutaneous and mesenteric lymph nodes), as were all
 6    pathological lesions identified on gross necropsy.
 7          There were no differences in food and water consumption or in body weights among the
 8    dose groups.  As illustrated graphically by the authors, there was little change in overall survival
 9    rates. Discussion of the histopathological findings focused exclusively on the cancer outcomes.
10    The incidence of total malignant tumors was increased significantly in high-dose males
11    compared to controls (p < 0.01). The slight increase in the incidence of total malignant tumors in
12    females was not  statistically significant (Table 4-29). With regard to the incidence of type- or
13    site-specific neoplasms, Soffritti et al. (2007) reported statistically significant increases
14    (calculated using the Cox regression model) in combined lymphomas and leukemias in both
15    sexes of Sprague-Dawley rats.  In males, the most frequently observed histiotypes were
16    lymphoblastic lymphomas involving the lung and mediastinal peripheral nodes, while in females,
17    the most commonly observed lesions were lymphocytic lymphomas and lympho-immunoblastic
18    lymphomas involving the thymus, spleen, lung and peripheral nodes. There was also an increase
19    in the incidence of mammary gland carcinomas in female Sprague-Dawley rats.  The incidences
20    of total malignant, mammary, and lymphocytic and leukocytic tumors, in comparison to
21    concurrent and the range of historical controls for combined lymphomas and leukemias and
22    mammary gland  tumors observed at the ERF, are shown in Table 4-29.

          Table 4-29. Incidence of tumors in  Sprague-Dawley rats exposed to aspartame from
          GD12 to natural death
Dose
(mg/kg-day)
Malignant tumors
Tumor-bearing
animals (percent)
Tumors/100
animals
Lymphomas/leukemias
Tumor-bearing animals
(percent)
Mammary carcinomas
Tumor-bearing animals
(percent)
Males
0
20
100
Historical controls
23/95 (24.2)
18/70 (25.7)
28/70 (40.0)a
ND
27.4
27.1
44.3
ND
9/95 (9.5)
11/70(15.7)
12/70 (17. l)b
8-31%
0/95 (0)
0/70 (0)
2/70 (2.9)
NR
Females
0
20
100
Historical controls
42/95 (44.2)
31/70(44.3)
37/70 (52.9)
ND
50.5
62.9
85.7
ND
12/95 (12.6)
12/70(17.1)
22/70 (33.4)a
7-18%
5/95 (5.3)
5/70(7.1)
1 1/70(1 5. 7)b
4-14%
                                              4-87
DRAFT-DO NOT CITE OR QUOTE

-------
Dose
(mg/kg-day)
Malignant tumors
Tumor-bearing
animals (percent)
Tumors/100
animals
Lymphomas/leukemias
Tumor-bearing animals
(percent)
Mammary carcinomas
Tumor-bearing animals
(percent)
1
2
3
4
5
6
7
      zp < 0.01 versus concurrent controls, as calculated by the authors using the Cox regression model.
      bp < 0.05 versus concurrent controls, as calculated by the authors using the Cox regression model.
      ND = no data.; NR = not reported.
      Source: Soffritti et al. (2007).

            With regard to target organs and tissues susceptible to aspartame carcinogenicity, Soffritti
      et al. (2007) drew attention to the similar outcome of these results to those reported by Soffritti
      et al. (2006, 2005).  The authors suggested that the increased incidence in combined lymphomas
      and leukemias in female Sprague-Dawley rats as compared to the earlier study was likely due to
      the earlier exposure to aspartame experienced by the animals in the Soffritti et al. (2007) study
      (prenatal and postnatal versus postnatal only). The authors provided a direct comparison of the
      incidence of lymphomas/leukemias between the studies, as summarized in Table 4-30.

          Table 4-30. Comparison of the incidence of combined lymphomas and leukemias in
          female Sprague-Dawley rats exposed to aspartame in feed for a lifetime, either pre-
          and postnatally or postnatally only
Dose (mg/kg-day)
0
20
100
Percent with lymphomas/leukemias
Pre-and postnatal exposure"
12.6
17.1
31.4
Postnatal exposure onlyb
8.7
20.0
18.7
 9
10
11
12
13
14
15
16
17
18
19
20
     Source: a Soffritti et al. (2007); b Soffritti et al. (2006, 2005).

           An EPS A (2009) review of the Sofftirti et al. (2007) study notes that the ratio between the
     incidence in the low-dose group and the incidence in the concurrent control in Table 4-30 is
     considerably lower in the animals exposed prenatally (1.4:1) compared to those exposed
     postnatally (2.3:1).  The ratio in the groups exposed to 100 mg/kg bw/day relative to the
     respective concurrent controls is only slightly higher in animals exposed prenatally (2.5:1)
     compared to those exposed postnatally (2.2:1).  EFSA also notes that the incidence of
     lymphomas and leukemias in aspartame-receiving males of the Soffritti et al. (2007) study was
     within the range of historical controls for these responses in Sprague-Dawley rats at the ERF.
     For example, the percent incidence of combined lymphomas and leukemias in males exposed
     pre- and postnatally to 100 mg/kg-day aspartame (17.1%) was within the range of historical
     controls for this response (8-31%, with an overall mean of 20.6%). Soffritti et al. (2007)
     acknowledge this fact, but reason that comparisons of potentially compound-associated
     incidences of tumor formation to incidences in concurrent controls provide a more scientifically
                                                           DRAFT-DO NOT CITE OR QUOTE

-------
 1    valid indicator of the tumorigenic impact of a chemical under investigation than comparisons to
 2    historical control data.57  Furthermore, the incidence of combined lymphomas and leukemias in
 3    the female rats in the high-dose group is well above the historical control range. Therefore,
 4    Soffritti et al. (2007) concluded that their second experiment confirmed the carcinogenic
 5    potential of aspartame in Sprague-Dawley rats observed in Soffritti et al. (2006, 2005).  The high
 6    and variable incidence of this tumor type in ERF controls remains a concern. However, the
 7    results provide support for studies suggesting similar effects from methanol (Soffritti et al.,
 8    2002a) since methanol is one of the degradation products of aspartame and appears to have
 9    carcinogenic potential at some of the same target organs and tissues.
10    4.6.5.2.3. MTBE.  In an experiment that also may be relevant to the carcinogenicity of
11    methanol, scientists at the ERF carried out a cancer bioassay on MTBE, in which the compound
12    was administered to 60 Sprague-Dawley rats/sex/group by  gavage in olive oil at 0, 250, and
13    1,000 mg/kg-day, 4 days/week, for 104 weeks (Belpoggi et al., 1995). Doses adjusted to daily
14    dose were 0, 143, and 571 mg/kg-day.  This experiment and its findings may relate to the
15    carcinogenicity of methanol, since methanol is one of several metabolites of MTBE (ATSDR,
16    1997). At the  end of the exposure period, the animals were allowed to live out their "natural"
17    life, the last animal dying 166 weeks after the start of the experiment (at 174 weeks of age).
18          Mean daily feed and drinking water consumption were determined weekly for the first
19    13 weeks of the experiment, then every 2 weeks until 112 weeks of age.  Individual body weights
20    were measured according to the same protocol, then every  8 weeks until the end of the
21    experiment. All animals were examined for gross lesions weekly for the first 13 weeks, then
22    every 2 weeks until term. All animals were examined grossly at death, then histopathologically
23    examined for a full suite of organs and tissues.
24          As described by the authors, there were no differences among the groups in body weight
25    and clinical signs of toxicity.  Survival was dose-dependently reduced in female rats after
26    16 weeks of exposure. Paradoxically, survival was improved in high-dose males compared to
27    controls after 80 weeks.  Although there were no noncarcinogenic effects of MTBE reported, a
28    number of benign and malignant  tumors were identified, including tumor types that were not
29    observed in the ERF methanol study such as an increased incidence of testicular Ley dig tumors
30    in high-dose males and as determined by the authors, as well as a dose-related statistically
31    significant increase in lymphomas and leukemias in females. The incidences of these tumors
32    compared to the initial number of animals exposed and compared to those  at risk at the time of
33    the first observed tumor formation are  shown in Table 4-31.
      57 There are also potential problems with the use of historical control information from a colony that has been
      maintained for over three decades. Population sensitivity can and does change over time.
                                               4-89        DRAFT-DO NOT CITE OR QUOTE

-------
          Table 4-31. Incidence of Leydig cell testicular tumors and combined lymphomas and
          leukemias in Sprague-Dawley rats exposed to MTBE via gavage for 104 weeks
Duration-
adjusted dose
0
143
571
Leydig cell tumors
Number of males
Affected
2
2
11"
Initial
60
60
60
At
riskc
26
25
32
Combined lymphomas and leukemias
Number of males
Affected
10
9
7
Initial
60
60
60
At
Riskd
59
59
58
Number of females
Affected
2
6b
12b
Initial
60
60
60
At
Risk6
58
51
47
      ap < 0.05 using prevalence analysis
      bp < 0.01 using a log-ranked test.
      0 Alive male rats at 96 weeks of age, when first Leydig cell tumor was observed.
      d Alive male rats at 32 weeks of age, when first leukemia was observed.
      e Alive female rats at 56 weeks of age, when first leukemia was observed.
      Source: Belpoggietal. (1995).
 1          The possible contribution of the metabolite methanol to the reported responses cannot be
 2    quantified. It is also possible that the parent compound and/or one or more of MTBE's other
 3    metabolites (e.g., tertiary butanol or formaldehyde) may be etiologically linked to the formation
 4    of the identified neoplasms (Blancato et al., 2007).
 5    4.6.5.2.4. Formaldehyde. Scientists at the ERF have carried out two long-term experiments on
 6    the potential carcinogenicity of formaldehyde, which is itself a metabolite of methanol,
 7    aspartame and MTBE. While the tumorigenic effects at the portal-of-entry (such as in the oral
 8    cavity and GI tract, for oral studies) may lack relevance to the possible effects of metabolites
 9    formed in situ following methanol exposure, systemic neoplasms such as lymphomas and
10    leukemias have been described for formaldehyde as well (Soffritti et al., 2002b, 1989). This
11    suggests that formaldehyde metabolized from methanol, aspartame and MTBE may be
12    etiologically important in the formation of lymphomas and leukemias in animals exposed to
13    these compounds.
14          In the first formaldehyde study (designated BT 7001; Soffritti et al., 1989), 50 Sprague-
15    Dawley rats/sex/group (starting at 7 weeks of age) were exposed to formaldehyde in drinking
16    water at concentrations of 10, 50, 500, 1,000, and 1,500 mg/L for 104 weeks.  Another 50
17    Sprague-Dawley rats/sex received methanol in  drinking water at 15 mg/L, and 100  rats/sex
18    received water only, as controls. Body weight and water and food consumption were monitored
19    weekly for the first 13 weeks, then every 2 weeks thereafter.  All animals were allowed to live
20    out their "natural" life, at which point they were subjected to necropsy and a complete
21    histopathological examination.
22          The final results of the BT 7001 Soffritti et al. (1989) experiment were reported by
23    Soffriti et al. (2002b).  Water consumption was reduced compared to controls in high-dose males
                                              4-90       DRAFT-DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
13
and in females at the three highest doses. However, there appeared to be no evidence of
compound-related body weight changes, clinical signs of toxicity among the groups, nor
nononcogenic histopathological effects of formaldehyde. The authors noted statistically
significant increases in the incidence of tumor-bearing males at 1500 ppm (p<0.01) and in total
malignant tumors in females at 100, 1000 and 1500 ppm (p<0.01) and in males at 500 ppm
(p<0.05) and 1500 ppm (p<0.01).  They reported statistically significant increases in malignant
mammary tumors in females at 100 ppm (p<0.01) and 1500 ppm (p<0.05), and in testicular
interstitial cell adenomas in the 1000 ppm males (p<0.05).  They also noted sporadic incidences
in the treatment groups only (primarily at the highest dose) of leiomyosarcomas of the stomach
and intestine considered to be very rare for the ERF rat colony.  As for methanol and the other
compounds discussed in this section, they reported increases in the number of
hemolymphoreticular tumors for both sexes. The incidence of hemolymphoreticular neoplasms
among the dose groups is shown in Table 4-32.

    Table 4-32.  Incidence of hemolymphoreticular neoplasms on Sprague-Dawley rats
    exposed to formaldehyde in drinking water for 104 weeks
Concentration in drinking water (mg/L)
0
0(1 5 mg/L methanol)
10
50
100
500
1,000
1,500
Males
8/100
10/50
4/50
10/50
13/50b
12/50a
ll/50a
23/50b
Females
7/100
5/50
5/50
7/50
8/50
7/50
ll/50a
10/50b
14
15
16
17
18
19
20
21
22
23
24
      "p < 0.05 using the tf test; bp < 0.01 using the x test.
      Source: Soffritti et al. (2002b).
       Soffritti et al. (1989) also described the results of another experiment (BT 7005) in which
approximately 20 Sprague-Dawley rats/sex/group were exposed to either regular drinking water
or 2,500 mg/L formaldehyde, beginning at 25 weeks of age for 104 weeks. These animals were
allowed to mate and approximately 40-60 of the FO pups were likewise exposed to 0 or
2,500 ppm formaldehyde in drinking water (after weaning) for 104 weeks. As before, parents
and progeny lived out their normal life span but then were subjected to a complete
histopathological examination. Incidence of leukemias in exposed breeders and offspring is
shown in Table 4-33.  The authors considered this data to indicate a "slight" increase in
leukemias in breeders at 2,500 ppm, but the changes did not achieve statistical significance.
                                               4-91
                                                     DRAFT-DO NOT CITE OR QUOTE

-------
          Table 4-33. Incidence of leukemias in breeder and offspring Sprague-Dawley rats
          exposed to formaldehyde in drinking water for 104 weeks (Test BT 7005)
Concentration (ppm)
0
2,500
Incidence of leukemias
Breeder
Males
0/20
2/18
Females
1/20
2/18
Offspring
Males
3/59
4/36
Females
3/49
0/37
     Source: Soffrittietal. (1989).
     4.7. SYNTHESIS OF MAJOR NONCANCER EFFECTS
     4.7.1. Summary of Key Studies in Methanol Toxicity
 1          A substantial body of information exists on the toxicological consequences to humans
 2   who consume or are acutely exposed to large amounts of methanol. Neurological and
 3   immunological effects have been noted in adult human subjects acutely exposed to as low as
 4   200 ppm (262 mg/m3) methanol (Mann et al., 2002; Chuwers et al., 1995). Nasal irritation
 5   effects have been reported by adult workers exposed to 459 ppm (601 mg/m3) methanol.  Frank
 6   effects such as blurred vision and bilateral or unilateral blindness, coma, convulsions/tremors,
 7   nausea, headache, abdominal pain, diminished motor skills, acidosis, and dyspnea begin to occur
 8   as blood levels approach 200 mg methanol/L, and 800 mg/L appears to be the threshold for
 9   lethality.  Data for subchronic, chronic or in utero human exposures are very limited.
10   Determinations regarding longer term effects of methanol are based primarily on animal studies.
11          An end-point-by-end-point survey of the primary effects of methanol in experimental
12   animals is given in the following paragraphs. Tabular summaries of the principal toxicological
13   studies that have examined the impacts of methanol when experimental animals were exposed to
14   methanol via the oral or inhalation routes are provided in Tables 4-34 and 4-35.  Most studies
15   focused on  developmental and reproductive effects. A large number of the available studies were
16   performed by routes of exposure (e.g., i.p.) that are less relevant to the assessment.  The data are
17   summarized separate sections that address oral exposure (Section 4.7.1.1) and inhalation
18   exposure (Section 4.7.1.2).

          Table  4-34. Summary of studies of methanol toxicity in experimental animals  (oral)
Species, strain,
number/sex
Rat
Sprague-Dawley
30/sex/group
Dose/duration
0, 100, 500, and
2,500 mg/kg-day for
13 wk
NOAEL
(mg/kg-day)
500
LOAEL
(mg/kg-day)
2,500
Effect
Reduction of brain
weights, increase in the
serum activity of ALT
and AP. Increased liver
weights
Reference
U.S. EPA
(1986c)
                                              4-92
DRAFT-DO NOT CITE OR QUOTE

-------
Species, strain,
number/sex
Rat
Sprague-Dawley
100/sex/group
Mouse
Swiss
Dose/duration
0, 500, 5,000, or
20,000 ppm (v/v) in
drinking water, for
104 wk. Doses were
approx. 0, 46.6, 466,
and 1,872 mg/kg-day
(male) and 0, 52.9,
529, and 2101 mg/kg-
day (female)
560, 1000 and 2 100
mg/kg/d (female)
and 550, 970, and
1800 mg/kg/d (male),
6 days/wk for life
NOAEL
(mg/kg-day)
466-529
ND
LOAEL
(mg/kg-day)
1,872-2,101
1,800-2,100
Effect
Increased incidence of
ear duct3 carcinomas,
lymphoreticular
tumors, and total
malignant tumors. No
noncancer effects
Increased incidence of
liver parenchymal cell
necrosis and malignant
lymphomas
Reference
Soffritti et al.
(2002a)
Apaja (1980)
Reproductive/developmental toxicity studies
Rat
Long-Evans
10 pregnant
females/group
Mouse
CD-I
8 pregnant
females and 4
controls
0 and 2,500 mg/kg-
day on either GD15-
GD17orGD17-
GD19.
4 g/kg-day in 2 daily
dosesonGD6-GD15
NA
NA
2,500
4,000
Neurobehavioral
deficits (such as
homing behavior,
suckling ability
Increased incidence of
totally resorbed litters,
cleft palate and
exencephaly. A
decrease in the number
of live fetuses/litter
Infurna and
Weiss (1986)
Rogers, et al.
(1993a)
NA = Not applicable; ND = Not determined; M= male, F=female.
aln an NTP evaluation of pathology slides from another bioassay from this laboratory in which a similar ear duct
carcinoma finding was reported (Soffritti et al., 2006, 2005), NTP pathologists interpreted a majority of these ear
duct responses as being hyperplastic, not carcinogenic, in nature (EFSA, 2006; Hailey, 2004).
     Table 4-35. Summary of studies of methanol toxicity in experimental animals
     (inhalation exposure)
Species, strain,
number/sex
Monkey
M. fascicularis,
lor 2
animals/group
Dog (2)
Rat
Sprague-Dawley
5 males/ group
Dose/duration
0, 3,000, 5,000,
7,000, or 10,000
ppm, 21 hr/day, for
up to 14 days
10,000 ppm for 3
min, 8 times/day for
100 days
0, 200, 2000, or
10,000 ppm, 8
hr/day, 5 days/wk for
up to 6 wk
NOAEL
(ppm)
ND
NA
NA
LOAEL
(ppm)
ND
NA
200
Effect
Clinical signs of toxicity, CNS
changes, including degeneration
of the bilateral putamen, caudate
nucleus, and claustrum. Edema
of cerebral white matter.
None
Transient reduction in plasma
testosterone levels
Reference
NEDO (1987)
Sayers et al.
(1944)
Cameron et al.
(1984)
                                              4-93
DRAFT-DO NOT CITE OR QUOTE

-------
Species, strain,
number/sex
Rat
Sprague-Dawley
5 males/ group
Rat
Sprague-Dawley
5/sex/group
Monkey
M. fascicularis
3/sex/group
Rat
Sprague-Dawley
10/sex/group
Rat
Sprague-Dawley
1 5/sex/group
Monkey
M. fascicularis
2 or 3 animals/
group/time point
Rat
F344
20/sex/group
Mouse
B6C3FJ
30/sex/group
Mouse
B6C3FJ
52-53/sex/group
Rat
F344
52/sex/group
Dose/duration
0, or 200 ppm,
6 hr/day, for either 1
or 7 days
0, 500, 2,000, or
5,000 ppm,
5 days/wk for 4 wk
0, 500, 2,000, or
5,000 ppm,
5 days/wk for 4 wk
0,300, or 3, 000 ppm,
6 hr/day, 5 days/wk
for 4 wk
0 or 2,500 ppm, 6
hr/day, 5 days/wk for
4wk
0, 10, 100, or
1,000 ppm, 21 hr/day
for either 7, 19, or 29
mo
0, 10, 100, or
1,000 ppm,
20 hr/day, for 12 mo
0, 10, 100, or
1000 ppm, 20 hr/day,
for 12 mo
0, 10, 100, or
1,000 ppm,
20 hr/day, for 12 mo
0, 10, 100, or
1,000 ppm,
-20 hr/day for 2 yr
NOAEL
(ppm)
NA
5,000
5,000
NA
NA
ND
ND
NA
NA
100
100
LOAEL
(ppm)
200
NA
NA
300
2,500
ND
ND
NA
NA
1,000
1,000
Effect
Transient reduction in plasma
testosterone levels
No compound-related effects
No compound-related effects
Reduction in size of thyroid
follicles
Reduction of relative spleen
weight in females,
histopathologic changes to the
liver, irritation of the upper
respiratory tract
Limited fibrosis of the liver
Possible myocardial and renal
effects
No compound-related effects
No clear-cut compound-related
effects
Increase in kidney weight,
decrease in testis and spleen
weights
Fluctuations in a number of
urinalysis, hematology, and
clinical chemistry parameters.
Development of pulmonary
adenoma/adenocarcinoma
(males), pheochromocytomas
(females)
Reference
Cameron et al.
(1985)
Andrews et al.
(1987)
Poon et al.
(1994)
Poon et al.
(1995)
NEDO (1987)
Reproductive/developmental toxicity studies
Rat
Sprague-Dawley
15/pregnant
females/group
Rat
Sprague-Dawley
36/pregnant
females/group
0, 5,000, 10,000, or
20,000 ppm, 7 hr/day
on either GDI-GDI 9
orGD7-GD15.
0, 200, 1,000, or
5000 ppm,
22.7 hr/day, on GD7-
GD17
5,000
1,000
10,000
5,000
Reduced fetal body weight,
increased incidence of visceral
and skeletal abnormalities,
including rudimentary and extra
cervical ribs
Late-term resorptions, reduced
fetal viability, increased
frequency of fetal
malformations, variations and
delayed ossifications.
Nelson et al.
(1985)
NEDO (1987)
4-94
DRAFT-DO NOT CITE OR QUOTE

-------
Species, strain,
number/sex
Rat
Sprague-Dawley
FI and F2
generations of a
two-generation
study
Rat
Sprague-Dawley
Follow-up study of
brain weights in FI
generation of 10-
14/sex/group in FI
generation
Mouse
CD-I
30-1 14 pregnant
females/group
Mouse
CD-I
12-17 pregnant
females/group
Rat
Long-Evans
6-7 pregnant
females/group
Rat
Long-Evans
10-12 pregnant
females/group
Monkey
M. fascicularis
12 monkeys/group
Dose/duration
0, 10, 100, or
1000 ppm, 20 hr/day;
FI- birth to end of
mating (M) or
weaning (F); F2-
birth to 8 wks
0, 500, 1,000, and
2,000 ppm; GDO
through F! generation
0, 1,000, 2,000,
5,000, 7,500, 10,000,
or 15,000 ppm,
7 hr/day on GD6-
GD15.
0 and 10,000 ppm on
two consecutive days
during GD6-GD 13
or on a single day
during GD5-GD9
0 or 15,000 ppm,
7 hr/day on GD7-
GD19
0 or 4,500 ppm from
GD10toPND21.
0, 200, 600, or
1800 ppm, 2.5
hr/day, 7 days/wk,
during premating,
mating and gestation
NOAEL
(ppm)
100
500
1,000
NA
NA
NA
ND
LOAEL
(ppm)
1,000
1,000
2,000
10,000
15,000
4,500
NDa
Effect
Reduced weight of brain,
pituitary, and thymus at 8, 16
and 24 wk postnatal in F! and at
8wkinF2
Reduced brain weight at 3 wk
and 6 wk (males only). Reduced
brain and cerebrum weight at 8
wk (males only)
Increased incidence of extra
cervical ribs, cleft palate,
exencephaly; reduced fetal
weight and pup survival,
Delayed ossification
Cleft palate, exencephaly,
skeletal malformations
Reduced pup weight
Subtle cognitive deficits
Shortened period of gestation;
may be related to exposure (no
dose-response),
neurotoxicological deficits
including reduced performance
in the VDR test; may be related
to premature births.
Reference

Rogers et al.
(1993a)
Rogers and
Mole (1997)
Stanton et al.
(1995)
Weiss et al.
(1996)
Burbacher
et al. (20004a,
2004b, 1999a,
1999b)
1
2
3
4
ND = Not determined due to study limitations such as small number of animals /time point/ exposure level
NA = Not applicable.
aGestation resulted in a shorter period of gestation in dams exposed to as low as 200 ppm (263 mg/m3). However,
because of uncertainties associated with these results, including clinical intervention and the lack of a dose-
response, EPA was not able to identify a definitive NOAEL or LOAEL from this study.

4.7.1.1. Oral
       There have been very few subchronic, chronic, or in utero experimental studies of oral
methanol toxicity. In one such experiment, an EPA-sponsored 90-day gavage study in Sprague-
Dawley rats suggested a possible effect of the compound on the liver (U.S. EPA, 1986c). In the
absence of gross or histopathologic evidence of toxicity, fluctuations on some clinical chemistry
                                             4-95        DRAFT-DO NOT CITE OR QUOTE

-------
 1    markers of liver biochemistry and increases in liver weights at the highest administered dose
 2    (2,500 mg/kg-day) justify the selection of the mid-dose level (500 mg/kg-day) as a NOAEL for
 3    this effect under the operative experimental conditions. That the bolus effect may have been
 4    important in the induction of those few effects that were apparent in the subchronic study is
 5    suggested by the outcome of lifetime drinking water study of methanol that was carried out in
 6    Sprague-Dawley rats by Soffritti et al. (2002a). According to the authors, no noncancer
 7    toxicological effects of methanol were observed at drinking water concentrations of up to
 8    20,000 ppm (v/v).  Based on default assumptions on drinking water consumption and body
 9    weight gain assumptions, the high concentration was equivalent to a dose of 1,780 mg/kg-day in
10    males and 2,177 mg/kg-day in females.  In the stated absence of any changes to parameters
11    reflective of liver toxicity in the Soffritti et al.  (2002a) study, the slight impacts to the liver
12    observed in the subchronic study at 2,500 mg/kg-day suggest the latter dose to be a minimal
13    LOAEL. Logically, the true but unknown threshold would at the high end of the range from 500
14    (the default NOAEL) to 2,500 mg/kg-day for liver toxicity via oral gavage.
15          Two studies have  pointed to the likelihood that oral exposure to methanol is associated
16    with developmental neurotoxicity or developmental deficits.  When Infurna and Weiss (1986)
17    exposed pregnant Long-Evans rats to 2% methanol in drinking water (providing a dose of
18    approximately 2,500 mg/kg-day), they observed no reproductive or developmental sequelae
19    other than from 2 tests within a battery of fetal behavioral tests (deficits in suckling ability  and
20    homing behavior). In the oral section of the Rogers et al.  (1993a) study, such  teratological
21    effects as cleft palate and exencephaly and skeletal malformations were observed in fetuses of
22    pregnant female mice exposed to daily gavage doses of 4,000 mg/kg methanol during GD6-
23    GDIS.  Likewise, an increase in totally resorbed litters and a decrease in the number of live
24    fetuses/litter appear likely to have been an effect of the compound. Similar skeletal
25    malformations were observed by Rogers and Mole (1997), Rogers et al. (1993a), and Nelson
26    et al.  (1985) following inhalation exposure.
      4.7.1.2. Inhalation
27          Some clinical signs, gross pathology, and histopathological effects of methanol have been
28    seen in experimental animals including adult nonhuman primates exposed to methanol vapor.
29    Results from an unpublished study (NEDO, 1987)  of M. fasciculoris monkeys, chronically
30    exposed to concentrations as low as 10 ppm for up to 29 months, resulted in histopathological
31    effects in the liver, kidney, brain and peripheral nervous system. These results were generally
32    reported as subtle and do not support a robust dose response over the range of exposure levels
33    used. Confidence in the methanol-induced findings of effects in adult nonhuman primates is
34    limited because this study utilized a small number (2-3) of animals/dose level/time of sacrifice
35    and inadequately reporting of results (i.e., lack of clear documentation of a concurrent control
                                               4-96       DRAFT-DO NOT CITE OR QUOTE

-------
 1    group). In addition, the monkeys used in this study were all wild-caught. All of these concerns
 2    limit the study's utility in derivation of an RfC.
 3          A number of studies have examined the potential toxicity of methanol to the male
 4    reproductive system (Lee et al., 1991; Cameron et al., 1985, 1984). The data from Cameron
 5    et al. (1985, 1984) showed a transient but not necessarily dose-related decrease in serum
 6    testosterone levels of male Sprague-Dawley rats. Lee et al. (1991) reported the appearance of
 7    testicular lesions in 18-month-old male Long-Evans rats that were exposed to methanol for
 8    13 weeks and maintained on folate-deficient diets. Taken together, the Lee et al. (1991) and
 9    Cameron et al. (1985,  1984) study results could indicate chemically-related strain on the rat
10    system as it attempts to maintain hormone homeostasis. However, the available data are
11    insufficient to definitively characterize methanol as a toxicant to the male reproductive system.
12          When Sprague-Dawley rats were exposed to methanol, 6 hours/day for 4 weeks, there
13    were some signs of irritation to the eyes and nose. Mild changes to the upper respiratory tract
14    were also described in Sprague-Dawley rats that were exposed for 4 weeks to up to 300 ppm
15    methanol (Poon et al., 1995).  Other possible effects of methanol in rats included a reduction in
16    size of thyroid follicles (Poon et al., 1994), panlobular vacuolation of the liver, and a decrease in
17    spleen weight (Poon et al., 1995).   NEDO (1987) reported dose-related increases in moderate
18    fatty degeneration in hepatocytes of male mice exposed via inhalation for 12 months, but this
19    finding was not observed in the NEDO (1987) 18-month mouse inhalation study.  Nodes
20    reported in the liver of mice from the  18-month study may have been precancerous, but the
21    18-month study  duration was not of sufficient duration to make a  determination.
22           One of the most definitive and quantifiable toxicological impacts of methanol when
23    administered to experimental animals via inhalation  is related to the induction of developmental
24    abnormalities in fetuses exposed to the compound in utero. Developmental effects have been
25    demonstrated in a number of species,  including monkeys, but particularly rats and mice. Most
26    developmental teratological effects appear to be more severe in the latter species. For example,
27    in the study of Rogers et al. (1993 a) in which pregnant female CD-I mice were exposed to
28    methanol vapors on GD6-GD15 at a range of concentrations, reproductive and fetal effects
29    included an increase in the number of resorbed litters, a reduction in the  number of live pups, and
30    increased incidence of exencephaly, cleft palate, and the number of cervical ribs.  While the
31    biological significance of the cervical rib effect has been the subject of much debate (See
32    discussion  of Chernoff and Rogers [2004] in Section 5), it appears to be  the most sensitive
33    indicator of developmental toxicity from this study, with a NOAEL of 1,000 ppm (1,310 mg/m3).
34    In rats, however, the most sensitive developmental effect, as reported in  the NEDO (1987) two-
35    generation inhalation studies, was a postnatal reduction in brain weight at 3, 6 and 8 weeks
36    postnatally, which was significantly lower than controls when pups and their dams were exposed

                                               4-97        DRAFT-DO NOT CITE OR QUOTE

-------
 1    to 1,000 ppm (1,310 mg/m3) during gestation and throughout lactation. The NOAEL reported in
 2    this study was 500 ppm (655 mg/m3).
 3          Rogers and Mole (1997) addressed the question of which period of gestation was most
 4    critical for the adverse developmental effects of methanol in CD-I rats.  Such malformations and
 5    anomalies as cleft palate, exencephaly, and a range of skeletal defects, appeared to be induced
 6    with a greater incidence when the dams were exposed on or around GD6.  These findings were
 7    taken to indicate that methanol is most toxic to embryos during gastrulation and in the early
 8    stages of organogenesis. However, NEDO (1987) gestation-only and two-generation studies
 9    showed that significant reductions in brain weight were observed at a lower exposure levels
10    when pups and their dams were exposed during lactation as well as gestation, indicating that
11    exposure during the later stages of organogenesis, including postnatal development, can
12    significantly contribute to the severity of the effects in this late-developing organ system.
13          In comparing the toxicity (NOAELs and LOAELs) for the onset of developmental effects
14    in mice and rats exposed in utero, there is suggestive evidence from the above studies that mice
15    may be more susceptible to methanol than rats. Supporting evidence for this proposition has
16    come from in vitro studies in which rat and mouse embryos were exposed to methanol in culture
17    (Andrews et al., 1993). Further evidence for species-by-species variations in the susceptibility of
18    experimental animals to methanol during organogenesis has come from experiments on monkeys
19    (Burbacher et al., 2004a, 2004b, 1999a,1999b). In these studies, exposure of monkeys to
20    methanol during premating, mating, and throughout gestation resulted in a shorter period of
21    gestation in dams exposed to as low as 200 ppm (263 mg/m3). The shortened gestation period
22    was largely the result of C-sections performed in the methanol-exposure groups "in response to
23    signs of possible difficulty in the maintenance of pregnancy," including vaginal bleeding.
24    Though  statistically significant, the finding of a shortened gestation length may be of limited
25    biological  significance. Gestational age, birth weight and infant size observations in all exposure
26    groups were within normal ranges for M. fascicularis monkeys, and vaginal bleeding 1-4 days
27    prior to delivery of a healthy infant does not necessarily imply a risk to the fetus (as cited in
28    CERHR, 2004).  An ultrasound examination could have substantiated fetal or placental problems
29    arising from presumptive pregnancy duress (see Section 4.3.2). As discussed in Section 4.4.2,
30    there is also evidence from this study that methanol  caused neurobehavioral effects in exposed
31    monkey infants that may be related to the  gestational exposure. However, the data are not
32    conclusive,and a dose-response trend is not robust. There is insufficient evidence to determine if
33    the primate fetus is more or less sensitive than rodents to methanol teratogenesis. Several other
34    uncertainties contributed to decreased confidence in the use of this primate in quantitative
35    estimates of risk.  These included: a mixture of wild- and colony-derived monkey mothers used
36    in the study; the use of a cohort design necessitated by the  complexity of this study also

                                                4-98        DRAFT-DO NOT CITE OR QUOTE

-------
 1    seemingly resulted in limitations in power to detect effects (e.g., Pagan test results for controls);
 2    and no apparent adjustment in statistical analysis for results from the neurobehavioral battery of
 3    tests employed leading to concern about inflation of type 1 error. Because of the uncertainties
 4    associated with these results, including the fact that the decrease in gestational length was not
 5    exacerbated with increasing methanol exposure, EPA was not able to identify a definitive
 6    NOAEL or LOAEL from this study. This study does support the weight of evidence for
 7    developmental neurotoxicity in the hazard characterization of low-level methanol exposure.
 8          Weiss et al. (1996) and Stanton et al. (1995) evaluated the developmental and
 9    developmental neurotoxicological effects of methanol exposure on pregnant female Long-Evans
10    rats and their progeny. In the former study, exposure of dams to 15,000 ppm (19,656 mg/m3),
11    7 hours/day on GD7-GD19 resulted in reduced weight gain in pups, but produced little other
12    evidence of adverse developmental  effects. The authors subjected the pups to a number of
13    neurobehavioral tests that gave little if any indication of compound-related changes.  This study,
14    while using high exposure levels, was limited in its power to detect effects due to the small
15    number of animals used.  In the Weiss et al. (1996) study, exposure of pregnant female Long-
16    Evans rats to 0 or 4,500 (0 and 5,897 mg/m3) methanol from GD6 to PND21 likewise provided
17    fluctuating and inconsistent results in a number of neurobehavioral tests that did not necessarily
18    indicate any compound-related impacts.  The finding of this study indicated subtle cognitive
19    defects not on the learning of an operant task but in the reversal learning.  This study also
20    reported exposure-related changes in neurodevelopmental markers of NCAMs on PND4.
21    NCAMs are a family of glycoproteins that is needed for migration, axonal outgrowth, and
22    establishment of the pattern for mature neuronal function.
23          Taking all of these findings into consideration reinforces the conclusion that the most
24    appropriate endpoints for use in the derivation of an RfC for methanol are associated with
25    developmental neurotoxicity and developmental toxicity.  Among an array of findings indicating
26    developmental neurotoxicity and developmental malformations and anomalies that have been
27    observed in the fetuses and pups of exposed dams, an increase in the incidence of cervical ribs of
28    gestationally exposed mice (Rogers et al., 1993a) and a decrease in the brain weights of
29    gestationally and lactationally exposed rats (NEDO, 1987) appear to be the most robust and most
30    sensitive effects.

      4.8. NONCANCER MO A INFORMATION
31          A review by Jacobsen and McMartin (1986) has provided a comprehensive summary of
32    the mechanism by which methanol brings about its acute toxic effects. Overwhelmingly, the
33    evidence points to methanol poisoning being a consequence of formate accumulation. This
34    compound is formed from formaldehyde under the action of ADH3. Formaldehyde itself is
                                               4-99       DRAFT-DO NOT CITE OR QUOTE

-------
 1    formed from methanol under the action of ADH1. Evidence for the involvement of formate
 2    comes from the delay in the onset of harmful symptoms, detection of formate in the blood
 3    stream, and the profound acidosis that develops 12-24 hours after exposure to methanol.
 4    Treatments for methanol poisoning include the i.v. administration of buffer to correct the
 5    acidosis, hemodialysis to remove methanol from the blood stream, and i.v. administration of
 6    either ethanol or fomepizole to inhibit the activity of ADH1. Therapies to increase endogenous
 7    levels of folate may enhance the activity of THF synthetase, an enzyme that catalyzes the
 8    oxidation of formate to CC>2. Jacobsen and McMartin (1986) have drawn attention to the
 9    accumulation of lactate in advanced stages of severe methanol poisoning, a possible consequence
10    of formate inhibition of mitochondrial respiration and tissue hypoxia. The additional decrease in
11    blood pH is likely to enhance the nonionic diffusion of formic acid across cell membranes, with
12    resulting CNS-depression, hypotension, and further lactate production.
13          Jacobsen and McMartin (1986) summarized a body of evidence that also points to the
14    formate-related acidosis as the etiologically important factor in ocular damage. The hypothesis
15    suggests that ocular toxicity is due to the inhibition of cytochrome oxidase in the optic nerve by
16    formate. This would cause inhibition of ATP formation and consequent disruption of optic nerve
17    function.
18          While it is well established that the toxic consequences of acute methanol poisoning arise
19    from the action of formate, there is less certainty on how the toxicological impacts of longer-
20    term exposure to lower levels of methanol are brought about. For example, since developmental
21    effects in experimental animals appear to be significant adverse effects associated with in utero
22    methanol exposure, it is important to determine potential MO As for how these specific effects
23    are brought about.
24          As described in Section 4.6.1, data from experiments carried out by Dorman et al. (1995),
25    formate is not the probable proximate teratogen in pregnant CD-I mice exposed to high
26    concentrations of methanol vapor. This conclusion is based on the fact that there appeared to be
27    little, if any, accumulation of formate in the blood of methanol-exposed mice, and exencephaly
28    did not occur until formate levels were grossly elevated. Another line of argument is based on
29    the observation that treatment of pregnant mice with a high oral dose of formate did not induce
30    neural tube closure defects at media concentrations comparable to those observed in uterine
31    decidual swelling after maternal exposure to methanol.  Lastly, methanol- but not formate-
32    induced neural tube closure defects in mouse embryos in vitro at media concentrations
33    comparable to the levels of methanol  detected in blood after a teratogenic exposure.
34          Harris et. al (Hansen et al., 2005; Harris et al., 2004, 2003) carried out a series of
35    physiological and biochemical experiments on mouse and rat embryos exposed to methanol,
36    formaldehyde and formate, concluding that the etiologically important substance for embryo

                                              4-100       DRAFT-DO NOT CITE OR QUOTE

-------
 1    dysmorphogenesis and embryolethality was likely to be formaldehyde rather than the parent
 2    compound or formate. Specific activities for enzymes involved in methanol metabolism were
 3    determined in rat and mouse embryos during the organogenesis period of 8-25 somites (Harris
 4    et al., 2003).  The experiment was based on the concept that differences in the metabolism of
 5    methanol to formaldehyde and formic acid by the enzymes ADH1, ADH3, and CAT may
 6    contribute to hypothesized differences in species sensitivity that were apparent in toxicological
 7    studies.  A key finding was that the activity of ADH3 (converting formaldehyde to formate) was
 8    lower in mouse VYS than that of rats throughout organogenesis, consistent with the greater
 9    sensitivity of the mouse to the developmental effects of methanol exposure. Another study
10    (Harris et al., 2004) which showed that the inhibition of GSH synthesis increases the
11    developmental toxicity of methanol also lends support to this hypothesis because ADH3-
12    mediated metabolism of formaldehyde is the only enzyme involved in methanol clearance that is
13    GSH-dependent. These findings provide inferential evidence for the proposition that
14    formaldehyde may be the ultimate teratogen through diminished ADH3 activity.  This concept is
15    further supported by the demonstration that the LOAELs for the embryotoxic effects of
16    formaldehyde in rat and mouse embryos were much lower than those for formate and methanol
17    (Hansen et al., 2005).  Taking findings from both sets of experiments together, Harris  et. al.
18    (Hansen et al., 2005; Harris et al., 2004, 2003) concluded that the demonstrable lower capacity of
19    mouse embryos to transform formaldehyde to formate (by ADH3) could explain the increased
20    susceptibility of mouse versus rat embryos to the toxic  effects of methanol.
21           While studies such as those by Harris et al. (2004, 2003) and Dorman et al. (1995,  1994)
22    strongly suggest that formate is not the  metabolite responsible for methanol's teratogenic effects,
23    there are still questions regarding the relative involvement of methanol versus formaldehyde. In
24    vitro evidence suggests that formaldehyde is the more embryotoxic moiety, but methanol would
25    likely play a prominent role, at least in terms of transport to the target tissue. The high reactivity
26    of formaldehyde would limit its unbound and unaltered transport as free formaldehyde from
27    maternal to fetal blood (Thrasher and Kilburn, 2001), and the capacity for the metabolism  of
28    methanol to formaldehyde is likely lower in the fetus and neonate versus adults (see discussion
29    in Section 3.3)
30           In humans, metabolism of methanol occurs primarily through ADH1, whereas in rodents
31    methanol clearance involves primarily CAT, as well as ADH1.  There are no known studies that
32    compare enzyme activities of human ADH1 and rodent CAT. Assuming that relative expression
33    and activity of ADH1 is comparable across species, rodents are expected to clear methanol more
34    rapidly than humans due to involvement of CAT.  In fact, even  among rodents the metabolism of
35    methanol may be quite different, as one study has demonstrated that the rate of methanol
36    oxidation in mice is twice the rate in rats, as well as nonhuman primates (Mannering et al.,

                                              4-101       DRAFT-DO NOT CITE OR QUOTE

-------
 1    1969). Despite a faster rate of methanol metabolism, mice have consistently shown higher blood
 2    methanol levels than rats following exposure to equivalent concentrations (Tables 3-4 and 3-5).
 3    A faster respiration rate and increased fraction of absorption by mice is thought to be the reason
 4    for the higher blood methanol levels compared to rats (Perkins et al., 1995a). Using the exposure
 5    conditions of Horton et al. (1992) for rats, when the respiration rate scaling coefficient (QPC)
 6    was increased from the rat value of 16.4 to the mouse value of 25.4 while holding all other
 7    parameters constant, peak blood concentrations were predicted by the PBPK model to increase
 8    by 1.4-fold at 200 ppm and 1.8-fold at 2,000 ppm (where metabolism is becoming saturated).
 9    Because smaller species generally have faster breathing rates than larger species (in the PBPK
10    model, the respiration rate/BW is 3 times slower in humans versus rats and almost 10 times
11    slower versus mice), humans would be expected to accumulate less methanol than rats or mice
12    inhaling equivalent concentrations and given the same metabolism rate. However, Horton et al.
13    (1992) measured a blood concentration in rats  exposed to 200 ppm methanol of about 3.7 mg/L
14    after 6 hours of exposure while Sedevic et al. (1981) measured around 5.5  mg/L in human
15    volunteers after 6 hours of exposure to 231 ppm. Correcting for the higher exposure, human
16    blood concentrations would be around 4.8 mg/L if exposed at 200 ppm. Simulations with the
17    mouse model predict a blood level of 5.7 mg/L after 6 hours of exposure to 200 ppm, only 20%
18    higher than this interpolated human value. Thus the slower inhalation rate in humans is offset by
19    the slower metabolic rate, leading to equivalent blood concentrations. (If the same rate of
20    metabolism/BW as mice is used, human blood concentrations are predicted to decrease by
21    approximately fivefold.). These differences  are considered in Section 5 for the characterization
22    of human and rodent PBPK models used for the derivation of human equivalent concentrations
23    (HECs).

      4.9. EVALUATION OF CARCINOGENICITY
      4.9.1. Summary of Overall Weight-of-Evidence
24          Under the current Guidelines for Carcinogen Risk Assessment (U.S. EPA 2005a, 2005b),
25    methanol is likely to be carcinogenic to humans by all routes of exposure based on dose-
26    dependent trends in multiple tumors in both sexes of two strains of rats, by inhalation and oral
27    routes of exposure and increases in malignant lymphoma in both sexes of Swiss mice by oral
28    exposure.  Specifically, EPA's analysis of the Soffritti et al. (2002a) lifespan study of Sprague-
29    Dawley rats exposed to methanol in drinking water for 104 weeks indicates a statistically
30    significant increase in the incidence of lymphoma58 in lung and other organs at the two highest
      58 Combining lymphoblastic lymphomas, lymphocytic lymphomas, lympho-immunoblastic lymphomas and/or
      lymphoblastic leukemias as malignant lymphomas but excluding myeloid leukemias, histocytic sarcomas and
      monocytic leukemia as tumors of different origin (Cruzan, 2009; Hailey, 2004; McConnell et al., 1986).
                                               4-102       DRAFT-DO NOT CITE OR QUOTE

-------
 1    doses for males and across all doses for females (Fisher's exact, p< 0.05) and a statistically
 2    significant increase in relatively rare hepatocellular carcinomas in males compared to historical
 3    controls (n=407)59 (Fisher's exact/? < 0.05 for all doses and/? < 0.01 for the high-dose group).
 4    Statistically significant increases in the incidence of malignant lymphomas relative to historical
 5    controls (Fisher's exact,/? < 0.05) have also been observed in another rodent species, Swiss
 6    mice, following similar mg/kg-day exposures to methanol in drinking water for life (Apaja,
 7    1980).  The only available chronic inhalation studies of methanol (NEDO, 1985/2008a, 2008b)
 8    reported slight but statistically significant tumor responses in F344 rats at 24 months, and no
 9    evidence of carcinogenicity in B6C3Fi mice at 18 months. EPA's analysis of the NEDO
10    (1985/2008b) inhalation study of F344 rats indicates a dose-response trend (Cochrane-Armitage
11   p< 0.05) and an increased incidence over concurrent controls at the high dose (Fisher's exact
12   p< 0.05) of pulmonary adenomas/adenocarcinomas in male rats. This analysis also indicates a
13    statistically significant dose-response trend (Cochrane-Armitage/? < 0.05) and a statistically
14    significant increased incidence over NTP historical controls at the high-dose (Fisher's exact
15   p< 0.05) of pheochromocytomas in female rats.
16           This WOE conclusion is supported by the results of other studies performed by ERF that
17    have shown tumorigenic responses similar to that of methanol in male and female Sprague-
18    Dawley rats exposed to formaldehyde (via drinking water), a metabolite of methanol, and to
19    aspartame (via feed) and MTBE (via olive oil gavage),  substances that hydrolyze to release
20    methanol and formaldehyde.  Confidence in the designation of methanol as a likely human
21    carcinogen is strengthened by the fact that methanol  is metabolized to formaldehyde, a chemical
22    that has been classified as carcinogenic to humans (group 1)  (IARC, 2004).  As discussed below
23    and in Section 5.4.3, there are uncertainties in the interpretation of these findings. All of the key
24    studies have design and reporting limitations.  EPA has reanalyzed the reported data from both
25    the ERF (Soffritti et al., 2002a) and NEDO (1987, 1985/2008a, 2008b) studies.  In reassessing
26    the ERF study data, EPA decided to combine only those lymphomas considered to have
27    originated from the same cell type. In the case of the NEDO data, the significance of the tumor
28    findings were incompletely reported in the original NEDO (1987) summary.  Hence, EPA used
29    translations of the original, detailed Japanese study reports provided by NEDO and the Methanol
30    Institute (NEDO, 1985/2008a,2008b) and reanalyzed the individual animal data.

      4.9.2. Synthesis of Human, Animal, and Other Supporting Evidence
31           Evidence of the carcinogenic potential of methanol arises from drinking water studies in
32    Sprague-Dawley rats (Soffritti et al., 2002a) and in Eppley Swiss Webster mice (Apaja, 1980),
      59 Obtained by combining control data from ERF studies of methanol, formaldehyde, aspartame, MTBE, and
      TAME.available from the ERF website at http://www.ramazzini.it/fondazione/foundation.aspX
                                               4-103       DRAFT-DO NOT CITE OR QUOTE

-------
 1    and an inhalation study in F344 rats (NEDO, 1985/2008b), with no information available in
 2    humans. As is described in Section 4.2.1.3 (Table 4-2), Soffritti et al. (2002a) reported a number
 3    of tumors in methanol-exposed Sprague-Dawley rats. EPA reanalyzed the tumor findings from
 4    this study using individual animal pathology available from the ERF website (see Section
 5    5.4.1.1).60 As indicated above, the increase in a relatively rare hepatocellular carcinoma in males
 6    compared to historical controls (Fisher's exact/? < 0.05 for all doses andp < 0.01 for the high-
 7    dose group) is potentially related to methanol dosing. A significant  increase in the incidence of
 8    ear duct carcinoma was also reported by Soffritti et al. (2002a). However, the high incidence for
 9    this tumor in controls of the Soffritti et al. (2002a) study relative to other studies of Sprague-
10    Dawley rats (Cruzan, 2009) and the results of an NTP evaluation of pathology slides from
11    another bioassay (EFSA, 2006; Hailey, 2004) raise questions about the ear duct pathological
12    determinations of Soffritti et al. (2002a).61
13           As is described in Section 4.2.1.3 (Table 4-3), Apaja  (1980) found an increase in
14    malignant lymphomas in mid-dose (p = 0.06) and high-dose (p < 0.05) female and mid-dose
15    (p < 0.05) male Eppley Swiss Webster mice exposed for life via drinking water. The lack of a
16    concurrent unexposed control data limit the confidence that can be placed on the relevance of the
17    increased lymphoma responses noted in this study. However, while controls were not
18    concurrent, they were from proximate (within 3 years) generations of the same mouse colony,
19    lymphomas were evaluated via the same classification criteria and, in the case of the Hinderer
20    (1979) controls, the histopathological analysis was performed by the same author (Apaja, 1980).
21    In addition, this is a late developing tumor, as noted by the author, suggesting the possibility of a
22    higher tumor response in the females of all exposure groups  had their survival not been
23    significantly lower than untreated historical controls (see dose-response  analysis in Appendix E).
24    Further, additional  support for these study results comes from the  fact that, as described above,
25    Soffritti et al. (2002a) subsequently reported an increase in lymphomas following similar levels
26    of mg/kg-day methanol drinking water doses to Sprague-Dawley rats, another rodent species
27    with a high spontaneous lymphoma rate.
28           Chronic inhalation bioassays have been conducted in monkeys, mice, and F344 rats
29    (NEDO, 1987, 1985/2008a; 1985/2008b). No exposure-related carcinogenic responses were
30    observed in the monkey or mouse studies.  As is described in Section 4.2.2.3, individual tumor
31    responses from the rat study were not significantly increased over concurrent controls, but the
      60 ERF provided the EPA with the detailed, individual animal data via reports available through their web portal
      (http://www.ramazzini.it/fondazione/foundation.aspX This allowed the EPA to combine lymphomas of similar
      histopathological origins and confirm the tumor incidences reported in the Soffriti et al.  (2002a) paper.
      61 In an NTP evaluation of pathology slides from another bioassay from this laboratory in which a similar ear duct
      carcinoma finding was reported (Soffritti et al., 2006, 2005), NTP pathologists interpreted a majority of these ear
      duct responses as being hyperplastic, not carcinogenic, in nature (EFSA, 2006; Hailey, 2004).
                                                4-104        DRAFT-DO NOT CITE OR QUOTE

-------
 1    response in the high-dose (1,000 ppm) group for pulmonary adenomas/adenocarcinomas in male
 2    rats was increased over concurrent controls (Fisher's exact/? < 0.05), and the dose-response for
 3    both pulmonary adenomas/adenocarcinomas in male rats and pheochromocytomas in female rats
 4    represent increasing trends (Cochran-Armitage trend test/? < 05). Further, the high-dose
 5    responses for both of these tumor types were elevated (p < 0.05) over historical control
 6    incidences within their respective sex and strain. As can be seen from Table 4-5, the severity and
 7    combined incidence of effects reported in the alveolar epithelium of male rat lungs (epithethial
 8    swelling, adenomatosis, pulmonary adenoma and pulmonary adenocarcinoma) and the adrenal
 9    glands of female rats (hyperplasia and pheochromocytoma) were increased over controls and
10    lower exposure groups. This pathology and the appearance of a rare  adenocarcinoma in the
11    high-dose group are suggestive of a progressive effect associated with methanol exposure.  The
12    increased pheochromocytoma response in female rats is considered to be potentially treatment
13    related because this is a historically rare tumor type for female F344 rats (NTP, 2007, 1999;
14    Haseman et al., 1998)62 and because, when viewed in conjunction with the increased medullary
15    hyperplasia observed in the mid-exposure (100 ppm) group females,  it is indicative of a
16    proliferative change with increasing methanol exposure.
17          Additional support for the designation of methanol as a likely carcinogen is provided by
18    the fact that methanol is metabolized to formaldehyde, which has been associated with increased
19    incidences of lymphoma and leukemia in humans (IARC, 2004). Furthermore, lymphomas
20    similar to those noted in Sprague-Dawley rats following exposure to  methanol in drinking water
21    and following a similar dose-response pattern were noted  in a bioassay for formaldehyde in
22    drinking water conducted by the same laboratory (Soffritti et al., 2002b,  1989) (Section 4.9.3).
23    These shared endpoints suggests that the  carcinogenic effects of methanol may result from its
24    conversion to formaldehyde, though the moiety and  MOA responsible for methanol-associated
25    tumor formation have not been identified.
26          Significant increases in the incidence of lymphoreticular tumors have also been reported
27    for other chemicals that convert in the body to methanol and/or formaldehyde including
28    aspartame  (Soffritti et al., 2007, 2006, 2005) and MTBE (Belpoggi et al., 1997, 1995).  In
29    contrast, no such tumors have been reported in a similar study conducted with a structurally
30    similar alcohol, ethanol (Soffritti et al., 2002a). In addition, epidemiological studies have
31    associated  formaldehyde exposure with increases in  the incidence of related
32    lymphohematopoietic tumors. While lymphomas are a rare finding in chronic laboratory
33    bioassays,  NCI (Hauptmann et al., 2003) and NIOSH (Pinkerton et al., 2004) have reported
34    increased lymphohematopoietic cancer risk, principally leukemia, in  humans from occupational
      ft1")
        Haseman et al. (1998) report rates for spontaneous pheochromocytomas in 2-year NTP bioassays of 5.7%
      (benign) and 0.3% (malignant) in male F344 rats and 0.3% (benign) and 0.1% (malignant) in female (n = 1517)
      F344 rats.
                                              4-105        DRAFT-DO NOT CITE OR QUOTE

-------
 1    exposure to formaldehyde.63 The similarities in tumor response across these chemicals, as well
 2    as a similar shape in the dose-response curve, supports the hypothesis that the common
 3    carcinogenic moiety for these compounds is the generation or presence of formaldehyde. The
 4    dose-response analysis discussed in Section 5 provides additional evidence supporting a role for
 5    the formaldehyde metabolite of methanol.  When "total metabolites in blood" predicted by a
 6    PBPK model was used as the dose metric, model fit to the dose-response data was significantly
 7    improved.
 8           As discussed in Section 4.6.5.2, there are challenges relative to the interpretation of the
 9    observed lymphoreticular tumors because there is no indication that ERF used specific pathogen-
10    free (SPF) rats (Schoeb et al., 2009), and the protocol for the studies conducted by the ERF
11    (Soffritti et al., 2002c) is different from 2-year bioassays conducted by NTP and NEDO.  A
12    distinct characteristic of the protocol for long-term bioassays conducted by the ERF is to
13    maintain animals until spontaneous death, rather than sacrificing them at the end of exposure at
14    104 weeks. This difference in  protocol may have an impact on the tumors observed compared to
15    a 2-year bioassay (Melnick et al., 2007).  The ERF methanol and ethanol studies (Soffritti et al.,
16    2002a), as well as the  aspartame studies (Soffritti et al., 2007, 2006) described in Section 4.6.5.2,
17    employed a large number of animals (100 or more per dose group) compared to a typical (e.g.,
18    NTP) cancer bioassay. In addition, the Sprague-Dawley rats used by ERF appear to have
19    increased sensitivity to certain  lymphoma responses relative to F344 rats that have been typically
20    used in NTP studies (Caldwell et al., 2008).64 According to Soffritti et al. (2007, 2006), the
21    overall incidence of lymphomas/leukemias in ERF  studies is 13.3% (range, 4.0-25.0%) in
22    female historical controls (2,274 rats) and 20.6% (range, 8.0-30.9%) in male historical controls
23    (2,265 rats). This background  rate is considered to be high relative to other tumor types and
24    relative to the background rate for this tumor type in Sprague-Dawley rats from other
25    laboratories (Cruzan, 2009; EFSA, 2006),65 However, it is in a range that can be considered
26    reasonable for studies that employ a large number of animals (Caldwell et al., 2009; Leakey et
27    al., 2003). These characteristics of ERF studies (i.e., lifetime observation, large number of
28    animals, and test strain sensitive to endpoint but with a relatively low control background rate
29    and mortality) may give them the sensitivity needed to detect a chemically related lymphoma
30    response.
      63IARC (2004) concluded that there was sufficient epidemiological evidence that formaldehyde causes
      nasopharyngeal cancer in humans but, also, that there was strong evidence for a causal association between
      formaldehyde and the development of leukemia in humans.
      64 F344 rats have a high mortality rate due to late-developing mononuclear cell leukemia, but the lymphoblastic and
      immunoblastic lymphomas reported in the Sprague-Dawley rat by ERF following methanol, MTBE, formaldehyde
      and aspartame administration are rarely diagnosed in the F344 rat (Caldwell et al., 2008).
      65 Cruzan (2009) reports that the incidences of total cancers derived from bloodforming cells, designated as
      hemolymphoreticular tumors by Ramazzini pathologists, is consistently about four times higher than the incidences
      of such tumors in SD rats recorded in the Charles River Laboratory  historical database (CRL database).
                                                4-106        DRAFT-DO NOT CITE OR QUOTE

-------
 1           Other aspects of ERF studies may impede their ability to reliably detect a chemically
 2    related response (EFSA, 2009, 2006). Chronic inflammatory responses have been reported in
 3    test animals of some ERF studies (EFSA, 2009, 2006), which may be the result of infections in
 4    test animals resulting from a bioassay design that does not employ SPF rats (Schoeb et al., 2009)
 5    and allows the rats to live out their "natural life span" in the absence of disease barriers (e.g.,
 6    fully enclosed cages).  In fact, the ERF has acknowledged that the primary cause of spontaneous
 7    death in their rats is respiratory infection (Caldwell et al., 2008; Ramazzini Foundation, 2006;
 8    Soffritti et al., 2006). Cruzan (2009) has suggested that respiratory infections in test animals of
 9    the Soffritti et al. (2002a) methanol study were not specific to older rats, as findings of lung
10    pathology were reported as often in rats dying prior to 18 months as in rats dying at or after
11    24 months.66
12           In their reviews of the recently published ERF studies on aspartame (Soffritti et al., 2007,
13    2006), the European Food Safety Authority (EFSA) have suggested that the increased incidence
14    of lymphomas/leukemias reported in treated rats was related to chronic respiratory disease in the
15    rat colony (EFSA, 2009, 2006), which they suggest was caused by aMycoplasmapulmonis
16    (M pulmonis) infection. EFSA felt that the increased incidence of these tumors was unrelated to
17    aspartame, given the high background incidence of chronic inflammatory changes such as
18    bronchopneumonia in the lungs of treated and untreated rats, and the concern that such tumors
19    might arise as a result of abundant lymphoid hyperplasia in the lungs of rats  suffering from
20    chronic respiratory disease.  The scientific evidence to support the EFSA opinion that
21    lymphomas/leukemias can result from chronic infection is limited (Schoeb et al., 2009; Caldwell
22    et al., 2008). Epithelial hyperplasias and lymphoid accumulations are commonly found in the
23    larynx and trachea of rats infected with M. pulmonis, but induction of lymphoma has not been
24    noted (Everitt and Richter, 1990; Lindsey et al., 1985). Further, the lung, not the larynx or
25    trachea, has been reported as the site of respiratory tract hemolymphoreticular tumors in ERF
26    studies of MTBE (Belpoggi et al., 1998, 1995) and methanol (Soffritti et al., 2002a).67 In their
27    review of the molecular biology and pathogenicity of M pulmonis, Razin et  al. (1998) note that
28    further study is needed before any conclusion can be reached regarding a relationship between
29    M. pulmonis and neoplasia.  In addition, if the increased incidence of lymphoreticular tumors in
30    the ERF methanol study was strictly the consequence of an incipient respiratory infection in the
31    ERF  rat colony, one would expect this to be a common finding across ERF studies. However, as
32    discussed in  Section 4.6.5.2, of the 200 compounds tested by ERF, only 8, which includes
      66 The infection rate did not have a significant impact on survival, however. The 2-year survival rate was 40-50% in
      the ERF methanol bioassay (see Appendix E, Figures E-l and E-2), which is above the average 2-year NTP study
      survival rate of 41.5% for Sprague-Dawley rats (Caldwell et al., 2008).
      67 ERF provided EPA with the detailed, individual animal data for the Soffriti et al. (2000a) via reports available
      through their web portal (http://www.ramazzini.it/fondazione/foundation.asp).
                                               4-107       DRAFT-DO NOT CITE OR QUOTE

-------
 1    methanol, have been associated with an increased incidence of hemolymphoreticular tumors.
 2    Further, the chemicals for which hemolymphoreticular tumors have been reported have chemical
 3    characteristics or physical properties in common,68 consistent with the hypothesis that the
 4    increased response is chemical-related.
 5          While evidence for a causal association between respiratory infections and lymphomas is
 6    limited, there is evidence that respiratory infections may have confounded the interpretation of
 7    lung lesions in the ERF studies.  Schoeb et al. (2009) state that lymphomas illustrated in two
 8    ERF studies (Figure 10 of Soffritti et al., 2005 and Figures 1-5 of Belpoggi et al., 1999) do not
 9    demonstrate the lymphoma type, cellular morphology, and organ distribution typical of
10    lymphoma in rats, but are consistent with "lymphocyte and plasma cell accumulation  in the lung
11    that is characteristic of M pulmonis disease." They suggest that, because M pulmonis disease
12    can be exacerbated by chemical treatment, a plausible alternative explanation for the dose-related
13    response reported in the MTBE, aspartame and methanol ERF bioassays is that the studies were
14    confounded by M. pulmonis disease and that lesions of the disease were interpreted as
15    lymphoma. However, several ERF lymphoma diagnoses in multiple rat organ systems, including
16    the lung, have been confirmed by an independent panel of six NIEHS pathologists (Hailey,
17    2004). Further, 60% of the lymphoma incidences reported in the ERF methanol study involved
18    organ systems other than the lungs (Schoeb et al., 2009).  The incidence of "lung-only"
19    lymphomas is evenly distributed across the control and dose groups of the methanol study such
20    that removing "lung-only" lympho-immunoblastic lymphomas from consideration (i.e., using
21    only lymphomas from other organ systems) does not significantly alter the dose-response for this
22    lesion (see Section 5.4.3.2).
23          Based on the NEDO (1987) summary report, IPCS (1997) concluded that "no  evidence of
24    carcinogenicity was found in either species [F344 rats and B6C3Fi mice]." This determination
25    was made based on Fisher's exact test results which indicated that the reported high-dose
26    pulmonary adenoma response in male rats and the high-dose pheochromocytoma response in
27    female rats were not statistically significant.  However, IPCS did not have translations of the
28    original NEDO mouse and rat chronic studies (NEDO, 1985/2008a; 1985/2008b), which
29    provided additional detail for EPA's analysis and reported combined lung adenoma and
30    adenocarcinoma results for high-dose male rats. In  addition, IPCS did not consider trend test
31    results or historical tumor data for F344 rats, both of which indicate a positive result for lung
32    adenoma/adenocarcinoma (males) and pheochromocytomas (females) from the NEDO rat study.
      68 Methanol, formaldehyde, aspartame, and MTBE, have common metabolites (e.g., formaldehyde); DIPE, TAME,
      methanol, and MTBE are all gasoline-oxygenate additives.
                                              4-108       DRAFT-DO NOT CITE OR QUOTE

-------
      4.9.3. MOA Information
 1          As discussed in Section 4.6.5.1, the results of genotoxicity/mutagenicity studies have
 2    been largely negative, irrespective of the presence or absence of metabolic activation (an S9
 3    microsomal fraction). Studies that investigate the MOA for methanol, particularly with respect
 4    to its developmental effects, have been discussed extensively in Sections 4.6. and 4.8.  Studies
 5    such as those by Harris et al. (2004, 2003) suggest that formaldehyde is the proximate teratogen
 6    and provide evidence in support of that hypothesis.  It is reasonable to hypothesize that the
 7    highly reactive molecule, formaldehyde, has a role in the carcinogenicity of methanol,  given the
 8    ability of formaldehyde to bind to proteins and DNA, induce DNA-protein cross-links, and
 9    possibly participate in reactions leading to free radical formation and the formation of lipid
10    peroxidation products.  As discussed in Section 4.6.3, evidence of oxidative stress following
11    methanol exposure has been reported in several organ  systems. Studies of Wistar rats suggest
12    that methanol  exposure can cause the production of free radical formation, lipid peroxidation,
13    and protein  modifications in the liver (Skrzydlewska et al.,  2005) and brain (Rajamani et al.,
14    2006), and adversely impact the oxidant/antioxidant balance in the brain (Dudka, 2006) and
15    lymphoid organs (Parthasarathy et al.,  2006b).
16          As discussed in Section 4.6.5.2, ERF studies of a number of compounds that have
17    formaldehyde as a metabolic product have been reported to cause lymphomas in
18    Sprague-Dawley rats. As described in Section 4.6.5.2.4, the ERF has conducted a formaldehyde
19    drinking water study (Soffritti et  al., 2002b, 1989) that is comparable in its design to the
20    methanol drinking water study of Soffritti et al. (2002a). The mg/kg-day doses of metabolized
21    methanol in Sprague-Dawley rats from the ERF methanol study estimated from the PBPK model
22    described in Section 3.4 and mg/kg-day doses of formaldehyde reported in the ERF
23    formaldehyde study were plotted together versus the hemolymphoreticular neoplasm incidences
24    in their respective studies (Figure 4-1).  Separate linear models were fit to the male and female
25    rat data from these studies.  The model fits shown in Figure 4-1 demonstrate that when
26    metabolized methanol is used as the dose metric for the methanol study data, the dose-response
27    data from these two studies can be adequately fit by two separate  linear dose-response functions
28    for the combined male (R2 = 0.6722) and combined female (R2 =  0.779) responses. Even if it is
29    true that formaldehyde is the common  moiety responsible for these tumors, one would not expect
30    this approach to result in perfect dose-response alignment because the metabolized methanol
31    estimate is not an accurate representation of formaldehyde distribution, and formaldehyde from
32    methanol administration would not be  expected to distribute the same as orally administered
33    formaldehyde. However, the similarities in the dose-response data for male and female rats from
34    these studies are consistent with the hypothesis that formaldehyde is key to methanol's
35    carcinogenic MOA.
                                               4-109       DRAFT-DO NOT CITE OR QUOTE

-------
                         0.5

                        0.45 -
                      I
                      tfl
                      <0  0.4
                      o.
                      o
                      0>
                      Z 0.35 -
                      J2
                      .1  0.3 -
                      •5 0.15-1
                      o>
                      !  c
                      s
                      o
                      — 0.05 -
  y = 0.0492Ln(x) + 0.0932
     R2 = 0.6832
                                          10.00               100.00
                                    mg formaldehyde/kg/d or mg metabolized methanol/kg/d
                             n Sofritti et al. (2002b), formaldehyde - males
                             • Sofiritti et al. (2002a), methanol - females
• Sof itti et al. (2002a), methanol - males
o Sofrritti et al. (2002b), formaldehyde - females
          Figure 4-1. Hemolymphoreticular neoplasms in male and female Sprague-Dawley rats
          in formaldehyde and methanol drinking water studies versus mg formaldehyde/kg/day
          or mg metabolized methanol/kg/day (predicted by EPA PBPK model).
            Source: Soffritti et al. (2002b).

 1          As discussed above, methanol is metabolized to formaldehyde, which is deemed to be
 2   carcinogenic to humans (IARC, 2004) by both the oral and inhalation routes, and there are
 3   readily apparent similarities between the dose-response data from oral studies of rats exposed to
 4   formaldehyde and methanol.  In addition, the dose-response model fit for the lymphoma
 5   response observed in the Soffritti et al. (2002a) study is improved when predicted total
 6   metabolites is used as the dose-metric (Section 5.4.1.2). However, the database of information
 7   available concerning methanol's carcinogenic MOA is limited, and the extent to which the parent
 8   or a metabolite such as formaldehyde is responsible for the carcinogenic effects observed in the
 9   studies conducted by Soffritti et al. (2002a) or NEDO (1987, 1985/2008b) is not clear.

     4.10. SUSCEPTIBLE POPULATIONS AND LIFE  STAGES
     4.10.1. Possible  Childhood Susceptibility
10          Studies in animals have identified the fetus as being more sensitive than adults to the
11   toxic effects of methanol; the greatest susceptibility occurs during gastrulation and early
12   organogenesis (CERHR, 2004). Table 4-21 summarizes some of the data regarding the relative
13   ontogeny  of CAT, ADH1, and ADH3 in humans and mice.  Human fetuses have limited ability to
14   metabolize methanol as ADH1 activity in 2-month-old and 4-5 month-old fetuses is 3-4% and
                                               4-110       DRAFT-DO NOT  CITE OR QUOTE

-------
 1    10% of adult activity, respectively (Pikkarainen and Raiha, 1967). ADH1 activity in 9-22 week
 2    old fetal livers was found to be 30% of adult activity (Smith et al., 1971).  Likewise, ADH1
 3    activity is -20-50% of adult activity during infancy (Smith et al., 1971; Pikkarainen and Raiha,
 4    1967). Activity continues to increase until reaching adult levels at 5 years of age (Pikkarainen
 5    and Raiha, 1967).  However, no difference between blood methanol levels in 1-year-old infants
 6    and adults was  observed following ingesting the same doses of aspartame, which releases 10%
 7    methanol by weight during metabolism (Stegink et al., 1983). Given that the exposure was
 8    aspartame as opposed to methanol, it is difficult to draw any conclusions from this study vis-a-
 9    vis ontogeny data and potential influences of age differences in aspartame disposition. With
10    regard to inhalation exposure, increased breathing rates relative to adults may result in higher
11    blood methanol levels in children compared to adults (CERHR, 2004). It is also possible that
12    metabolic variations resulting in increased methanol blood levels in pregnant women could
13    increase the fetus'  risk from exposure to methanol. In all, unresolved issues regarding the
14    identification of the toxic moiety increase the uncertainty with regards to the extent and
15    pathologic basis for early life susceptibility to methanol  exposure.
16          The prevalence of folic acid deficiency has decreased since the United States and Canada
17    introduced a mandatory folic acid food fortification program in November 1998. However,
18    folate deficiency is still a concern among pregnant and lactating women, and factors such as
19    smoking, a poor quality diet, alcohol intake, and folic antagonist medications can enhance
20    deficiency (CERHR, 2004). Folate deficiency could affect a pregnant woman's ability to clear
21    formate, which has also been demonstrated to produce developmental toxicity in rodent in in
22    vitro studies at  high-doses (Dorman et al., 1995). It is not known if folate-deficient humans have
23    higher levels of blood formate than individuals with adequate folate levels. A limited study  in
24    folate-deficient monkeys demonstrated no formate accumulation following an endotracheal
25    exposure of anesthetized monkeys to 900  ppm methanol for 2 hours (Dorman et al., 1994).  The
26    situation is obscured by the fact that folic  acid deficiency during pregnancy by itself is thought to
27    contribute to the development of severe congenital malformations (Pitkin, 2007).

      4.10.2. Possible Gender Differences
28          There is limited information on potential differences in susceptibility to the toxic effects
29    of methanol according to gender. However, one study reported a higher background blood
30    methanol level  in human females versus males (Batterman and Franzblau, 1997).  In rodents,
31    fetuses exposed in utero were found to be the most sensitive subpopulation.  One study suggested
32    a possible increased sensitivity of male versus female rat fetuses and pups. When rats were
33    exposed to methanol pre- and postnatally, 6- and 8-week-old male progeny had significantly
34    lower brain weights at 1,000 ppm, compared to those in  females that demonstrated the same
                                               4-111       DRAFT-DO NOT CITE OR QUOTE

-------
 1   effect only at 2,000 ppm (NEDO, 1987). In general, there is little evidence for substantial
 2   disparity in the level or degree of toxic response to methanol in male versus female experimental
 3   animals or humans. However, it is possible that the compound-related deficits in fetal brain
 4   weight that were evident in the pups of FI generation Sprague-Dawley rats exposed to methanol
 5   in the NEDO (1987) study may reflect a threshold neurotoxicological response to methanol.  It is
 6   currently unknown whether higher levels of exposure would result in brain sequelae comparable
 7   to those observed in acutely exposed humans.

     4.10.3.  Genetic Susceptibility
 8          Polymorphisms in enzymes involved in methanol metabolism may affect the sensitivity
 9   of some individuals to methanol. For example, as discussed in Chapter 3, data summarized in
10   reviews by Agarwal (2001), Burnell  et al. (1989), Bosron and Li (1986), and Pietruszko (1980)
11   discuss genetic polymorphisms for ADH. Class IADH, the primary ADH in human liver, is a
12   dimer composed of randomly associated polypeptide units encoded by three genetic loci
13   (ADH1 A, ADH1B, and ADH1C). Polymorphisms are observed at the ADH1B (ADH1B*2,
14   ADH1B*3) and ADH1C (ADH1C*2) loci.  The ADH1B*2 phenotype is estimated to occur in
15   -15% of Caucasians of European descent, 85% of Asians, and less that 5% of African
16   Americans. Fifteen percent of African Americans have the ADH1B*3 phenotype, while it is
17   found in less than 5% of Caucasian Europeans and Asians.  The only reported polymorphisms in
18   ADH3 occur in the promoter region, one of which reduces the transcriptional activity in vitro
19   nearly twofold (Hedberg et al., 2001).  While polymorphisms in ADH3 are described in more
20   than one report (Cichoz-Lach et al., 2007; Hedberg et al., 2001), the functional consequence(s)
21   for these polymorphisms remains unclear.
                                             4-112       DRAFT-DO NOT CITE OR QUOTE

-------
                 5. DOSE-RESPONSE ASSESSMENT AND CHARACTERIZATION

      5.1. INHALATION RfC FOR EFFECTS OTHER THAN CANCER69
 1           In general, the RfC is an estimate of a daily exposure of the human population (including
 2    susceptible subgroups) that is likely to be without an appreciable risk of adverse health effects
 3    over a lifetime. It is derived from a POD, generally the statistical lower confidence limit on the
 4    BMCL or BMDL, with uncertainty/variability factors applied to reflect limitations of the data
 5    used. The inhalation RfC considers toxic effects for both the respiratory system (portal-of-entry)
 6    effects  and systems peripheral to the respiratory system (extra-respiratory  or systemic effects). It
 7    is generally expressed in mg/m3.  EPA performed an IRIS assessment of methanol in  1991  and
 8    determined that the database was inadequate for derivation of an RfC. While some limitations
 9    still exist in the database (see  Sections 5.1.3.2 and 5.3), the experimental toxicity database has
10    expanded and newer methods and models have been developed to analyze the results. In this
11    update, the PBPK model, described in Section 3.4, was developed by EPA and is used to estimate
12    HECs and HEDs from inhalation study data for the derivation of both the RfC and RfD. In both
13    cases, the use of a PBPK model replaces part of the UF adjustments traditionally used for
14    species-to-species extrapolation.
15           Additionally, this assessment uses the BMD method in its derivation of the POD.70 The
16    suitability of these methods to derive a POD is dependent on the nature of the toxicity database
17    for a specific chemical. Details of the BMD analyses are found in Appendix C. The  use of the
18    BMD approach for determining the POD improves the assessment by including consideration of
19    shape of the dose-response curve,  independence from experimental doses, and estimation of the
20    uncertainty pertaining to the calculated dose response.  However, the methanol database still has
21    limitations and uncertainties associated with it, in particular, those uncertainties associated with
22    human variability, animal-to-human differences, and limitations in the database influence
23    derivation of the RfC.

      5.1.1. Choice of Principal Study and Critical Effect(s)
      5.1.1.1. Key Inhalation Studies
24           While a substantial body of information exists on the toxicological consequences to
25    humans exposed to large amounts of methanol, no human studies exist that would allow for
      69 The RfC discussion precedes the RfD discussion in this assessment because the inhalation database ultimately
      serves as the basis for the RfD. The RfD development would be difficult to follow without prior discussion of
      inhalation database and PK models used for the route-to-route extrapolation.
      70 Use of BMD methods involves fitting mathematical models to dose-response data and using the results to select a
      POD that is associated with a predetermined benchmark response (BMR), such as a 10% increase in the incidence of
      a particular lesion or a 10% decrease in body weight gain (see Section 5.1.2.2).
                                                 5-1       DRAFT—DO NOT CITE OR QUOTE

-------
 1    quantification of sub chronic, chronic, or in utero effects of methanol exposure. Table 4-35
 2    summarizes available experimental animal inhalation studies of methanol.  Several of these
 3    studies, including the monkey chronic (NEDO, 1987) and developmental (Burbacher et al.,
 4    2004a, 2004b, 1999a, 1999b) studies, the male rat reproductive studies (Lee et al., 1991;
 5    Cameron et al., 1985, 1984), and the 4-week rat studies (Poon et al., 1994), are lacking in key
 6    attributes (e.g., documented dose response, documented controls, and duration of exposure)
 7    necessary for their direct use in the quantification of a chronic RfC.  These studies will be
 8    considered in this chapter for their contributions to the overall RfC uncertainty.  Several
 9    inhalation reproductive or developmental studies were adequately documented and are of
10    appropriate size and design for quantification and derivation of an RfC.  These studies are
11    considered for use in the derivation of an RfC and are summarized below.
      5.1.1.2. Selection of Critical Effect(s)
12          Developmental effects have been assessed in a number of toxicological studies of
13    monkeys, rats, and mice. The findings of Rogers and Mole (1997) indicate that methanol is toxic
14    to mouse embryos in the early stages of organogenesis, on or around GD7. In the study of
15    Rogers et al. (1993a), in which pregnant female CD-I mice were exposed to methanol vapors
16    (1,000, 2,000, and 5,000  ppm) on GD6-GD15, reproductive and fetal effects included an
17    increase in the number of resorbed litters, a reduction in the number of live pups, and increased
18    incidences of exencephaly, cleft palate, and the number of cervical ribs.  They reported a
19    NOAEL for cervical rib malformations at 1,000 ppm (1,310 mg/m3) and a LOAEL of 2,000 ppm
20    (2,620 mg/m3, 49.6% per litter versus 28.0% per litter in the control group). Increased incidence
21    of cervical ribs was also observed in the rat organogenesis study (NEDO, 1987)  in the 5,000 ppm
22    dose group (65.2% per litter versus 0% in the control group), indicating that the  endpoint is
23    significant across species.
24          The biological significance of the cervical rib endpoint within the regulatory arena has
25    been the subject of much debate (Chernoff and Rogers, 2004).  Previous studies  have classified
26    this endpoint as either a malformation (birth defect of major importance) or a variation
27    (morphological alternation of minor significance). There is evidence that incidence of
28    supernumerary ribs (including cervical ribs) is not just the addition of extraneous, single ribs but
29    rather is related to a general alteration in the development and architecture of the axial skeleton
30    as a whole.  In CD-I mice exposed during gestation to various types of stress, food and water
31    deprivation, and the herbicide dinoseb, supernumerary ribs were consistently associated with
32    increases in length of the 13th rib (Branch et al., 1996). This relationship was present in all fetal
33    ages examined in the  study.  The authors concluded that these findings are consistent with
34    supernumerary ribs being one manifestation of a basic alteration in the differentiation of the
35    thoraco-lumbar border of the axial skeleton. The biological significance of this endpoint is
36    further strengthened by the association of supernumerary ribs with adverse health effects in
                                                5-2       DRAFT—DO NOT CITE OR QUOTE

-------
 1    humans. The most common effect produced by the presence of cervical ribs is thoracic outlet
 2    disease (Nguyen et al., 1997; Fernandez Noda et al., 1996; Henderson, 1914).  Thoracic outlet
 3    disease is characterized by numbness and/or pain in the shoulder, arm, or hands. Vascular effects
 4    associated with this syndrome include cerebral and distal embolism (Beam et al., 1993; Connell
 5    et al., 1980; Short, 1975), while neurological symptoms include extreme pain, migraine, and
 6    symptoms similar to Parkinson's (Evans, 1999; Saxton et al., 1999; Fernandez Noda et al., 1996).
 7    Schumacher et al. (1992) observed 242 rib anomalies in 218 children with tumors (21.8%) and
 8    11 (5.5%) in children without malignancy, a statistically significant (p < 0.001) difference that
 9    indicates a strong association between the presence of cervical ribs and childhood cancers.
10    Specific cancers associated with statistically significant increases in anomalous ribs included
11    leukemia, brain, tumor, neuroblastoma, soft tissue sarcoma, and Wilm's tumor.
12          A number of rat studies have confirmed the toxicity of methanol to embryos during
13    organogenesis (Weiss et al.,  1996; NEDO, 1987; Nelson et al., 1985).  NEDO (1987) reported
14    reduced brain, pituitary, and thymus weights in FI and F2 generation Sprague-Dawley rats at
15    1,000 ppm methanol.  In a follow-up study of the FI generation brain weight effects, NEDO
16    (1987) reported decreased brain, cerebellum, and cerebrum weights in FI males exposed at
17    1,000 ppm methanol from GDO through the FI generation.  The exposure levels used in these
18    studies are difficult to interpret because dams were exposed prior to gestation, and dams and
19    pups were exposed during gestation and lactation. However, it is clear that postnatal exposure
20    increases the severity of brain weight reduction.  In another experiment in which NEDO (1987)
21    exposed rats only during organogenesis (GD7-GD17), the observed decreases in brain weights in
22    offspring at 8 weeks of age were less severe than in the studies for which exposure was
23    continued postnatally. This finding is not unexpected, given that the brain undergoes tremendous
24    growth beginning early in gestation and continuing in the postnatal period. Rats are considered
25    altricial (i.e., born at relatively underdeveloped stages), and many of their neurogenic events
26    occur postnatally (Clancy et al., 2007). Brain effects from postnatal exposure are also relevant to
27    humans given that, in humans, gross measures of brain growth increase for at least 2-3 years
28    after birth, with the growth rate  peaking approximately 4 months after birth (Rice and Barone,
29    2000).
30          A change in brain weight is considered to be a biologically significant effect (U.S. EPA,
31    1998). This is true regardless of changes in body weight because brain weight is generally
32    protected during malnutrition or weight loss, unlike many other organs or tissues
33    (U.S. EPA,  1998). Thus, change in absolute brain weight is an appropriate measure of effects on
34    this critical  organ system. Decreases in brain weight have been associated with simultaneous
35    deficits in neurobehavioral and cognitive parameters in animals exposed during gestation to
36    various solvents, including toluene and ethanol (Gibson, 2000; Coleman et al., 1999; Hass,

                                                5-3       DRAFT—DO NOT CITE OR QUOTE

-------
 1    1995). NEDO (1987) reports that brain, cerebellum, and cerebrum weights decrease in a dose-
 2    dependant manner in male rats exposed to methanol throughout gestation and the FI generation.
 3          Developmental neurobehavioral effects associated with methanol inhalation exposure
 4    have been investigated in monkeys.  Burbacher et al. (2004a, 2004b, 1999a,  1999b) exposedM
 5   fascicularis monkeys to 0, 262, 786, and 2,359 mg/m3 methanol, 2.5 hours/day, 7 days/week
 6    during premating/mating and throughout gestation (approximately 168 days). In these studies,
 7    exposure of monkeys to up to 1,800 ppm (2,359 mg/m3) methanol during premating, mating, and
 8    throughout gestation resulted in no changes in reproductive parameters other than a shorter
 9    period of gestation in all exposure groups that did not appear to be dose related.  The shortened
10    gestation period was largely the result of C-sections performed in the methanol exposure groups
11    "in response to signs of possible difficulty in the maintenance of pregnancy," including vaginal
12    bleeding. As discussed in  Section 4.7.1.2, though statistically significant, the shortened gestation
13    finding may be of limited biological significance given questions concerning its relation to the
14    methanol exposure. Developmental parameters, such as fetal crown-rump length and head
15    circumference, were unaffected, but there appeared to be neurotoxicological deficits in  methanol -
16    exposed pups. VDR was significantly reduced in the 786 mg/m3 group for males and the 2,359
17    mg/m3 group for both sexes. However,  a dose-response trend for this endpoint was only
18    exhibited for females.  In fact, this is the only effect reported in the Burbacher et al. (2004a,
19    2004b, 1999a, 1999b) studies for which a significant dose-response trend is evident. As
20    discussed in Section 4.4.2, confidence may have been increased by statistical analyses to adjust
21    for multiple testing (CERHR, 2004). Yet it is worth noting that the dose-response trend for VDR
22    in females remained significant with (p = 0.009) and without (p = 0.0265) an adjustment for the
23    shortened gestational periods, and it is a measure of functional deficits in sensorimotor
24    development that is consistent with early developmental CNS effects (brain weight changes
25    discussed above) that have been observed in rats.
26          Another test, the Fagan test of infant intelligence, indicated small but not significant
27    deficits of performance (time spent looking at novel faces versus familiar faces) in treated
28    monkeys. Although not statistically significant and not  quantifiable, the results of this test are
29    also important when considered in conjunction with the brain weight changes noted in the NEDO
30    (1987) rat study.  As discussed in Section 4.7.1.2, the monkey data are not conclusive, and there
31    is insufficient evidence to determine if the primate fetus is more or less sensitive than rodents to
32    methanol teratogenesis. Taken together, however, the NEDO (1987) rat study and the Burbacher
33    et al. (2004a, 2004b, 1999a,  1999b) monkey study suggest that prenatal exposure to methanol
34    can result in adverse effects on developmental neurology pathology and function, which can be
35    exacerbated by continued postnatal exposure.
36          A number of studies described in Section 4.3.2 and summarized in Section 4.7.1.2 have
37    examined the potential toxicity of methanol to the male reproductive system (Lee et al., 1991;
                                                5-4       DRAFT—DO NOT CITE OR QUOTE

-------
 1   Cameron et al., 1985, 1984).  Some of the observed effects, including a transient decrease in
 2   testosterone levels, could be the result of chemically related strain on the rat system as it attempts
 3   to maintain hormone homeostasis.  However, the data are insufficient to definitively characterize
 4   methanol as a toxicant to the male reproductive system.
 5          The studies considered for use in the derivation of an RfC are summarized in Table 5-1.
 6   As discussed in Sections 5.1.3.1 and 5.3, there is uncertainty associated with the selection of an
 7   effect endpoint from the methanol database for use in the derivation of an RfC. Taking into
 8   account the limitations of the studies available for quantification purposes, decreased brain
 9   weight at 6 weeks in male Sprague-Dawley rats  exposed throughout gestation and the postnatal
10   period (NEDO, 1987) was chosen as the critical  effect for the purposes of this dose-response
11   assessment as it can be reliably quantified and represents both a sensitive organ system and a key
12   period of development.  RfC derivations utilizing alternative endpoints (e.g., cervical rib effects
13   in mice and delayed sensorimotor development in monkeys) and alternative methods (e.g., use of
14   different BMRLs) are summarized in Appendix C and in Section 5.1.3.1.

          Table 5-1. Summary of studies considered most appropriate for use in derivation of an
          RfC
Reference
NEDO (1987)
Two-generation
study
NEDO (1987)
Follow-up study
of FI generation
NEDO (1987)
Teratology study
Nelson et al.
(1985)
Species
(strain)
Rat
Sprague-
Dawley
Rat
Sprague-
Dawley
Rat
Sprague-
Dawley
Sex
M,F
M,F
F
Number/
dose group
Not specified -
F! and F2
generation
10-147 sex/
group- F!
generation
10-12/sex/
group
15 pregnant
dams/group
Exposure
Duration
Fj- Birth to end
of mating (M)
or weaning (F);
F2-birth to 8
wk
GDO through
F! generation
GD7-GD17
GDI-GDI 9 or
GD7-GD15
Critical Effect
Reduced weight of
brain, pituitary, and
thymus at 8, 16, and
24 wk postnatal in F!
and at 8 wk in F2
Reduced brain weight
at 3 wk and 6 wk
(males only).
Reduced brain and
cerebrum weight at 8
wk (males only)
Reduced brain,
pituitary, thyroid,
thymus, and testis
weights at 8 wk
postnatal.
Reduced fetal body
weight, increased
incidence of visceral
and skeletal
abnormalities,
including rudimentary
and extra cervical ribs
NOAEL
(ppm)
100
500
1,000
5,000
LOAEL
(ppm)
1,000
1,000
5,000
10,000
                                                5-5
DRAFT—DO NOT CITE OR QUOTE

-------
Reference
Rogers et al.
(1993a)
Burbacher et al.
(20004a, 2004b,
1999a, 1999b)
Species
(strain)
Mouse
CD-I
M.
fascicularis
Sex
F

Number/
dose group
30-114
pregnant dams/
group
12 pregnant
monkeys/group
Exposure
Duration
GD6-15
2.5 hr/day,
7 days/wk,
during
premating,
mating and
gestation
Critical Effect
Increased incidence
of extra cervical ribs,
cleft palate,
exencephaly; reduced
fetal weight and pup
survival, delayed
ossification
Shortened period of
gestation; may be
related to exposure
(no dose response),
neurotox. deficits
including reduced
performance in the
VDR test
NOAEL
(ppm)
1,000

LOAEL
(ppm)
2,000
b
      "Animals were dosed 20-21 hr/day. NS = Not Specified
      bGestational exposure resulted in a shorter period of gestation in dams exposed to as low as 200 ppm (263 mg/m3).
      However, because of uncertainties associated with these results, including clinical intervention and the lack of a
      dose-response, EPA was not able to identify a definitive NOAEL or LOAEL from this study.

      5.1.2. Methods of Analysis for the POD—Application of PBPK and BMD Models
 1          Potential PODs for the RfC derivation, described here and in Appendix C, have been
 2    calculated via the use of monkey, rat and mouse PBPK models, described in Section 3.4.  First,
 3    the doses  used in an experimental bioassay were converted to an internal dose metric that is most
 4    appropriate for the endpoint being assessed.  The PBPK models are capable of calculating
 5    several measures of dose for methanol, including the following:
 6           •  Cmax- The peak concentration of methanol in the blood during the exposure period;
 7           *A UC - Area under the curve, which represents the cumulative product of
 8             concentration and time for methanol in the blood; and
 9           •  Total metabolism - The production of metabolites of methanol, namely formaldehyde
10             and formate.
11          As described in Section 3.4, the focus of model development is on obtaining accurate
12    predictions of increased body burdens over background.  The PBPK models do not  describe or
13    account for background levels of methanol, formaldehyde or formate.
14          Although there remains uncertainty surrounding the identification of the proximate
15    teratogen  of importance (methanol, formaldehyde, or formate), the dose metric  chosen for
16    derivation of an RfC was based on blood methanol levels. This decision was primarily based on
17    evidence that the toxic moiety is not likely to be the formate metabolite of methanol (CERHR,
18    2004) and evidence that levels of the formaldehyde metabolite following methanol maternal
19    and/or neonate exposure would be much lower in the fetus and neonate than in adults. While
20    recent in vitro evidence indicates that formaldehyde is more embryotoxic than methanol and
21    formate, the high reactivity of formaldehyde would limit its unbound and unaltered transport as
                                               5-6       DRAFT—DO NOT CITE OR QUOTE

-------
 1    free formaldehyde from maternal to fetal blood (Thrasher and Kilburn, 2001), and the capacity
 2    for the metabolism of methanol to formaldehyde is likely lower in the fetus and neonate versus
 3    adults (see discussion in Section 3.3).  Thus, even if formaldehyde is identified as the proximate
 4    teratogen, methanol would likely play a prominent role, at least in terms of transport to the target
 5    tissue. Further discussions of methanol metabolism, dose metric selection, and MOA issues are
 6    covered in Sections 3.3, 4.6, 4.8 and 4.9.2.
 7          A BMDL was then derived in terms of the internal dose metric utilized.  Finally, the
 8    BMDL values were converted to HECs via the use of a PBPK model parameterized for humans.
 9    The next section describes the rationale for and application of the benchmark modeling
10    methodology for the RfC derivation.
      5.1.2.1. Application of the BMD/BMDL Approach
11          Several developments over the past few years impact the derivation of the RfC: 1) EPA
12    has developed draft BMD assessment methods (U.S. EPA, 2000b, 1995) and supporting software
13    (Appendix C) to improve upon the previous NOAEL/LOAEL approach; 2) MOA studies have
14    been carried out that can give more insight into methanol toxicity; and 3) EPA has refined PBPK
15    models for methanol on the basis of the work of Ward et al. (1997) (see Section 3.4. for
16    description of the EPA model).  The EPA PBPK model provides estimates of HECs from rodent
17    exposures that are supported by pharmacokinetic information available for rodents and humans.
18    The following sections describe how the BMD/BMDL approach, along with the EPA PBPK
19    model, is used to obtain a POD for use in the derivation of an RfC for methanol in accordance
20    with current draft BMD technical guidance (U.S. EPA, 2000b).
21          The BMD approach attempts to fit models to the dose-response data for a given endpoint.
22    It has the advantage of taking more of the dose-response data into account when determining the
23    POD, as well as estimating the dose for which an effect may  have a specific probability of
24    occurring. The BMD approach also accounts, in part, for the quality of the study (e.g., study
25    size) by estimating a BMDL, the 95% lower bound confidence limit on the BMD.  The BMDL is
26    closer to the BMD (higher) for large studies and further away from the BMD (lower) for small
27    studies. Because the BMDL approach will account, in part, for a study's power, dose spacing,
28    and the steepness of the dose-response curve, it is generally preferred over the NOAEL
29    approach.
30          When possible, all experimental data points are included in this assessment to ensure
31    adequate fit of a BMD model and derivation of a BMDL. A  summary of the POD values
32    determined by BMD analysis for the critical endpoint (as well as other considered endpoints)
33    (see Appendix C for modeling results), application of UFs, and conversion to HECs using the
34    BMD and PBPK approach, is included in Section 5.1.3.1.
                                               5-7       DRAFT—DO NOT CITE OR QUOTE

-------
 1           Use of the BMD approach has uncertainty associated with it. An element of the BMD
 2    approach is the use of several models to determine which best fits the data.71  In the absence of
 3    an established MOA or a theoretical basis for why one model should be used over another, model
 4    selection is based on best fit to the experimental data selection. Model fit was determined by
 5    statistics (AIC and ^ residuals of individual dose groups) and visual inspection recommended by
 6    EPA (2000b)72
 7           The PBPK model developed by EPA for methanol (described in Section 3.4) was applied
 8    for the estimation of methanol blood levels in the exposed dams (NEDO, 1987). When using
 9    PBPK models, it is very important to determine what estimate of internal dose (i.e., dose metric)
10    can serve as the most appropriate dose metric for the health effects under consideration.
11           The results of NEDO (1987), shown in Table 4-8, indicate that there is not a cumulative
12    effect of ongoing exposure on brain-weight decrements in rats exposed postnatally; i.e., the dose
13    response in terms of percent of control is about the same at 3 weeks postnatal as at 8 weeks
14    postnatal in rats exposed throughout gestation and the FI generation. However, there does
15    appear to be a greater brain-weight effect in rats exposed postnatally versus rats exposed only
16    during organogenesis (GD7-GD17).  In male rats exposed during organogenesis only, there is no
17    statistically significant decrease in brain weight at 8  weeks after birth at the 1,000 ppm exposure
18    level.  Conversely, in male rats exposed to the same  level of methanol throughout gestation and
19    the FI generation, there was an approximately 5% decrease in brain weights (statistically
20    significant at the/? < 0.01 level). The fact that male  rats exposed to 5,000 ppm methanol only
21    during organogenesis experienced a decrease in brain weight of 10% at 8 weeks postnatal
22    indicates that postnatal exposure is not necessary for the observation of persistent postnatal
23    effects.  However, the fact that this decrease was  less than the  13% decrease observed in male
24    rats exposed to 2,000 ppm methanol throughout gestation and the 8 week postnatal period
25    indicates that both exposure concentration and duration are important components of the ability
26    of methanol to cause this effect.  The extent to which the observation of the increased effect is
27    due to a cumulative effect in rats exposed postnatally versus recovery in rats for which exposure
28    was discontinued at birth is not clear.
29           The fact that brain weight is susceptible to both the level and duration of exposure
30    suggests that a dose metric that incorporates a time component would be the most appropriate
31    metric to use. For these reasons, and because it is more typically used in internal-dose-based
      71USEPA's BMDS 2.1 was used for this assessment as it provides data management tools for running multiple
      models on the same dose-response data set. At this time, BMDS offers over 30 different models that are appropriate
      for the analysis of dichotomous, continuous, nested dichotomous and time-dependent lexicological data. Results
      from all models include a reiteration of the model formula and model run options chosen by the user, goodness-of-fit
      information, the BMD, and the estimate of the lower-bound confidence limit on the BMD (BMDL).
      72Akaike's Information Criterion (AIC) (Akaike, 1973) is used for model selection and is defined as -2L + 2P where
      L is the log-likelihood at the maximum likelihood estimates for the parameters and P is the number of model degrees
      of freedom.
                                                 5-8        DRAFT—DO NOT CITE OR QUOTE

-------
 1    assessments and better reflects total exposure within a given day, daily AUC (measured for
 2    22 hours exposure/day) was chosen as the most appropriate dose metric for modeling the effects
 3    of methanol exposure on brain weights in rats exposed throughout gestation and continuing into
 4    the FI generation.
 5          Application of the EPA methanol PBPK model (described in Section 3.4) to the NEDO
 6    (1987) study, in which developing rats were exposed during gestation and the postnatal period,
 7    presents complications that need to be discussed.  The neonatal rats in this study were exposed to
 8    methanol  gestationally before parturition as well as lactationally and inhalationally after
 9    parturition.  The PBPK model developed by EPA only estimates internal dose metrics for
10    methanol  exposure in NP adult mice and rats. Experimental data indicate that inhalation-route
11    blood methanol kinetics in NP mice and pregnant mice on GD6-GD10 are similar (Dorman
12    et al., 1995; Perkins et al., 1995a,1995b; Rogers et al., 1993a,  1993b). In addition, experimental
13    data indicate that the maternal blood:fetal partition coefficient for mice is approximately 1 (see
14    Section 3.4.1.2). Assuming that these findings apply for rats, the data indicate that PBPK
15    estimates  of PK and blood dose metrics for NP rats are better predictors of fetal exposure during
16    gestation than would be obtained from default extrapolations from external exposure
17    concentrations. However, as is  discussed to a greater extent in Section 5.3, the additional routes
18    of exposure presented to the pups in this  study (lactation and inhalation) present uncertainties
19    that suggest the average blood levels in pups in the NEDO (1987) report might be greater than
20    those of the dam. The assumption made in this assessment is that, if such differences exist
21    between human mothers and their offspring, they are not expected to be significantly greater than
22    that which has been postulated for rats. Thus, the PBPK model-estimated adult blood methanol
23    level is considered to be an appropriate dose metric for the purpose of this analysis and HEC
24    derivation.
      5.1.2.2. BMD Approach Applied to Brain Weight Data in Rats
25          The NEDO (1987) study reported decreases in brain weights in developing rats exposed
26    during gestation only (GD7-GD17) or during gestation and the postnatal period,  up to 8 weeks.
27    Because of the biological significance of decreases in brain weight as an endpoint in the
28    developing rat and because this endpoint was not evaluated in other peer-reviewed studies, BMD
29    analysis was performed using these data. For the purposes of deriving an RfC for methanol from
30    developmental endpoints using the BMD method and rat data, decreases in brain weight at
31    6 weeks of age in the more sensitive gender, males, exposed throughout gestation and continuing
32    into the FI generation (both through lactation and inhalation routes) were utilized. Decreases in
33    brain weight at 6 weeks (gestation and postnatal exposure), rather than those seen at 3 and 8
34    weeks, were chosen as the basis for the RfC derivation because they resulted in lower estimated
35    BMDs and BMDLs. Decreased brain weights in male rats at 8 weeks age after gestation-only

                                                5-9       DRAFT—DO NOT CITE OR QUOTE

-------
 1    exposure were not utilized because they were less severe at the same dose level (1,000 ppm)
 2    compared to gestation and postnatal exposure.
 3          The first step in the current BMD analysis is to convert the inhalation doses, given
 4    as ppm values from the studies, to an internal dose metric using the EPA PBPK model (see
 5    Section 3.4). For decreased brain weight in male rats, AUC of methanol in blood (hr x mg/L) is
 6    chosen as the appropriate internal dose metric for the reasons discussed in Section 5.1.2.1.
 7    Predicted AUC values for methanol in the blood of rats are summarized in Table  5-2. These
 8    AUC values are then used as the  dose metric for the BMD analysis of decreased brain weight in
 9    male rats.73 The full details of this analysis are reported in Appendix C. More details concerning
10    the PBPK modeling were presented in Section 3.4.

          Table 5-2. The EPA PBPK model estimates of methanol blood levels (AUC) in rats
          following inhalation exposures
Exposure level (ppm)
500
1,000
2,000
Methanol in blood AUC
(hr x mg/L)a in Rats
79.2
226.7
967.8
      aAUC values were obtained by simulating 22 hr/day exposures for 5 days and calculated for the last 24 hours of that
      period.
11          The current draft BMD technical guidance (U.S. EPA, 2000b) suggests that, in the
12    absence of knowledge as to what level of response to consider adverse, a change in the mean
13    equal to one S.D. from the control mean can be used as a BMR for continuous endpoints.
14    However, it has been suggested that other BMRs, such as 5% change relative to estimated
15    control mean, are also appropriate when performing BMD analyses on fetal weight change as a
16    developmental endpoint (Kavlock et al., 1995). Therefore, both a one S.D. change from the
17    control mean and a 5% change relative to estimated control mean were considered (see Appendix
18    C for RfC derivations using alternative BMRs). For this endpoint, a one S.D. change from the
19    control mean returned the lowest BMDL estimates and was considered the most suitable BMR
20    for use in the RfC derivation. All models were fit using restrictions and option settings
21    suggested in the draft EPA BMD technical guidance document (U.S. EPA, 2000b).
22          A summary of the results most relevant to the development of a POD using the BMD
23    approach (BMD, BMDL, and model fit statistics) for decreased brain weight at 6 weeks in male
24    rats exposed to methanol throughout gestation and continuing into the FI generation is provided
25    in Table 5-3.  BMDL values in Table 5-3 represent the 95% lower-bound confidence limit on the
26    AUC estimated to result in a mean that is one S.D. from the control mean. There is a 2.5-fold
       All BMD assessments in this review were performed using BMDS version 2.1. A copy of the BMDS can be
      obtained at: .
                                               5-10      DRAFT—DO NOT CITE OR QUOTE

-------
 1    range of BMDL estimates from adequately fitting models, indicating considerable model
 2    dependence.  In addition, the fit of the Hill and more complex Exponential models is better than
 3    the other models in the dose region of interest as indicated by a lower scaled residual at the dose
 4    group closest to the BMD (0.09 versus -0.67 or -0.77) and visual inspection.  In accordance with
 5    draft EPA BMD Technical Guidance (EPA, 2000b), the BMDL from the Hill model (bolded), is
 6    selected as the most approriate basis for an RfC derivation because it results in the lowest BMDL
 7    from among a broad range of BMDLs and provides a superior fit in the low dose region nearest
 8    the BMD. The Hill model dose-response curve for decreased brain weight in male rats is
 9    presented in Figure 5-1, with response plotted against the chosen internal dose metric of AUC of
10    methanol in rats. The BMDLiso was determined to be 90.9 hr x mg/L using the 95% lower
11    confidence limit of the dose-response curve expressed in terms of the AUC for methanol in
12    blood.

          Table 5-3. Comparison of benchmark dose modeling results for decreased brain weight
          in male rats at 6 weeks of age using modeled AUC of methanol as a dose metric
Model
Linear
2nd degree polynomial
3rd degree polynomial
Power
Hillb
Exponential 2
Exponential 3
Exponential 4
Exponential 5
BMD1SD (AUC,
hr x mg/L) a
278.30
278.30
278.30
278.30
170.57
260.94
260.94
172.08
172.08
BMDL1SD
(AUC,
hr x mg/L)a
225.30
225.30
225.30
225.30
90.93
209.10
209.10
96.93
96.93
/7-value
0.5376
0.5376
0.5376
0.5376
0.8366
0.612
0.612
0.8205
0.8205
AICC
-203.84
-203.84
-203.84
-203.84
-203.04
-204.10
-204.10
-203.10
-203.10
Scaled residual"1
-0.77
-0.77
-0.77
-0.77
0.09
-0.67
-0.67
0.09
0.09
      "The BMDL is the 95% lower confidence limit on the AUC estimated to decrease brain weight by 1 control mean
      S.D. using BMDS model options and restrictions suggested by EPA BMD technical guidance (U.S. EPA, 2000b).
      bln accordance with draft EPA BMD Technical Guidance (EPA, 2000), the BMDL from the Hill model (bolded) is
      chosen for us in an RfC derivation because it is the lowest of a broad range of BMDL estimates from adequately
      fitting models and because the Hill model provides good fit in the dose region of interest as indicated by a relatively
      low scaled residual at the dose group closest to the BMD (0.09 versus -0.67 or -0.77).
      °AIC = Akaike Information Criterion = -2L + 2P, where L is the log-likelihood at the maximum likelihood
      estimates for the parameters, and P is the number of modeled degrees of freedom (usually the number of parameters
      estimated).
      Vd residual (measure of how model-predicted responses deviate from the actual data) for the dose group closest to
      the BMD scaled by an estimate of its S.D. Provides a comparative measure of model fit near the BMD. Residuals
      that exceed 2.0 in absolute value should cause one to question model fit in this region.

      Source: NEDO(1987).
                                                  5-11
DRAFT—DO NOT CITE OR QUOTE

-------
                                         Hill Model with 0.95 Confidence Level
                  8-
                  0)
                  or
                        1.85
                         1.8
                        1.75
                         1.7
1.65
                         1.6
                        1.55
                         1.5
                        1.45
                   09:05 08/24 2009
                                      200
                                               400      600
                                                  dose
                                                                800
                                                                         1000
         Figure 5-1. Hill model BMD plot of decreased brain weight in male rats at 6 weeks age
         using modeled AUC of methanol in blood as the dose metric, 1 control mean S.D.
 1
 2          Once the BMDLiso was obtained in units of hr x mg/L, it was used to derive a chronic
 3   RfC. The first step is to calculate the HEC using the PBPK model described in Appendix B. An
 4   algebraic equation is provided (Equation 1 of Appendix B) that describes the relationship
 5   between predicted methanol AUC and the human equivalent inhalation exposure concentration
 6   (HEC) in ppm.

 7          BMDLHEC (ppm) = 0.0224*BMDLlSD+(1334*BMDLlSD)/(794+ BMDL1SD)
 8               BMDLHEC (ppm) = 0.0224*90.9+( 1334*90.9)7(794+ 90.9) =  139 ppm

 9          Next, because RfCs are typically expressed in units of mg/m3, the HEC value in ppm was
10   converted using the conversion factor specific to methanol of 1 ppm =1.31 mg/m3:
11
   HEC (mg/m3) = 1.31 x 139 ppm = 182 mg/m3
                                             5-12
                               DRAFT—DO NOT CITE OR QUOTE

-------
      5.1.3. RfC Derivation - Including Application of Uncertainty Factors
      5.1.3.1. Comparison Bet ween En dp oin ts an d BMDL Modeling Approa ch es
 1          A summary of the PODs for the various developmental endpoints and BMD modeling
 2    approaches considered for the derivation of an RfC, along with the UFs applied74 and the
 3    conversion to an HEC, are presented in Table 5-4 and graphically compared in Figure 5-2 (see
 4    Appendix C for details). Information is presented that compares the use of different endpoints
 5    (i.e., cervical rib, decreased brain weight, and increased latency  of VDR) and different methods
 6    (i.e., different BMR levels) for estimating the POD. These comparisons are presented to inform
 7    the analysis of uncertainty surrounding these choices. Each approach considered for the
 8    determination of the POD has strengths and limitations, but when considered together for
 9    comparative purposes they allow for a more informed determination for the POD for the
10    methanol RfC.
11          A 10% extra risk BMR is adequate for most traditional bioassays using 50 animals per
12    dose group. A smaller BMR of 5% extra risk can sometimes be justified for developmental
13    studies (e.g., Rogers et al., 1993a), because they generally involve a larger number of subjects.
14    Reference values estimated for cervical rib incidence in mice using Cmax as the dose metric were
15    13.6 and 10.4 mg/m3 using BMDLio and BMDLos PODs, respectively (see Appendix D for
16    discussion of choice of Cmax as the appropriate dose metric for incidence of cervical rib in mice).
17    The reference value estimated for alterations in sensorimotor development and performance as
18    measured by the VDR test in female monkeys using AUC as the dose metric was 1.7 mg/m3
19    using the BMDLSo as the POD. As discussed in Section 4.4.2, confidence in this endpoint is
20    reduced by a marginal dose-response trend in one sex (females)  and a limited sample size.
21    Although the VDR test demonstrates that prenatal and continuing postnatal exposure to methanol
22    can result in neurotoxicity, the use of such statistically borderline results is not warranted in the
23    derivation  of the RfC, given the availability of better dose-response data in other species.
24    Decreases in brain weight at 6 weeks of age in male rats exposed during gestation and
25    throughout the FI generation using AUC as the dose metric yield the reference values of 1.8 and
26    2.4 mg/m3 for BMRs of one S.D. from the control mean and 5% change relative to control mean,
27    respectively. Because decreases in brain weight in male rats at 6 weeks postbirth resulted in a
28    clear dose response and returned RfC estimates lower than or approximate to the other endpoints
29    considered, it was chosen as the critical endpoint. One S.D. from the  control mean was chosen
30    as the appropriate level of response (BMR) for the  calculation of the RfC because it is the
31    standard recommended by EPA's draft technical guidance (U.S. EPA,  2000b) and yields a lower
32    BMDL than 5% relative deviance for this data set.  Thus, the RfC is:
      74 The rationale for the selection of these UFs is discussed later in Section 5.1.3.
                                               5-13       DRAFT—DO NOT CITE OR QUOTE

-------
         RfC = PODHEc - UF = 182 mg/m3 + 100 = 2 mg/m3 (rounded to one significant figure)
     Table 5-4. Summary of PODs for critical endpoints, application of UFs and conversion
     to HEC values using BMD and PBPK modeling


BMDL
HEC (mg/m3)c
UFHd
UFAe
UFD
UFS
UFL
UF-TOTAL
RfC (mg/m3)
Rogers et al. (1993a)
BMDL10 mouse
cervical rib Cmax
94.3 mg/L
1360
10
3
3
1
1
100
13.6
BMD L05 mouse
cervical rib Cmax
44.7 mg/L
1036
10
3
3
1
1
100
10.4
Burbacher et al.
(1999a,1999b)
BMDL1SD
female monkey
VDRa AUC
81.7 hrxmg/L
165
10
3
3
1
1
100
1.7
NEDO (1987)
BMDLos
rat brain wt. b
AUC
123. 9 hrxmg/L
240
10
3
3
1
1
100
2.4
BMDL1SD
rat brain wt. b
AUC
90.9 hrxmg/L
182
10
3
3
1
1
100
1.8
aVDR = test of sensorimotor development as measured by age from birth at achievement of test criterion for
grasping a brightly colored object.
bBrain weight at 6 weeks postbirth, multiple routes of exposure (whole gestation, lactation, inhalation)
cThe PBPK model used for this HEC estimate is described in Appendix B. An algebraic equation (Equation 1 of
Appendix B) describes the relationship between predicted methanol AUC and the human equivalent inhalation
exposure concentration (HEC) in ppm. This equation can also be used to estimate model predictions for HECs
from  Cmax values because Cmax values and AUC values were estimated at steady-state for constant 24 hours
exposures (i.e., AUC = 24 xCmax).  The ppm HEC estimate is then converted to mg/m3 by multiplying by 1.31.
dThe rationale for the selection of these UFs is discussed in Section 5.1.3 below.
These uncertainty factor (UF) acronyms are defined in Sections 5.1.2.1.1 to 5.1.2.1.4.
                                               5-14
DRAFT—DO NOT CITE OR QUOTE

-------
           10000
            1000
         £   100
                 Rogers etal., 1993a;GD Rogers et al., 1993a; GD Burbacher et al., 1999a,b; NEDO, 1987; 1-gen repro NEDO, 1987; 1-gen repro
                   6-15 mouse study;     6-15 mouse study;   whole gestation monkey study; 6 week male brain study; 6 week male brain
                     cervical rib;     cervical rib BMDL(HEC)05 study; visually directed   weight; BWDL(HEC)05  weight; BMDL(HEC)1SD
                    BMDL(HEC)10                     reaching; BMDL(HEC)1SD
                     POD   I  I Interspecies Extrapolation
Inter-individual Variation   I   I Database
                                     RfC
          Figure 5-2. PODs (in mg/m3) for selected endpoints with corresponding applied UFs
             (chosen RfC value is circled)
      5.1.3.2. Applica tion of UFs
 1            UFs are applied to the POD, identified from the rodent data, to account for recognized
 2    uncertainties in extrapolation from experimental conditions to the assumed human scenario (i.e.,
 3    chronic exposure over a lifetime). A composite UF of 100-fold (10-fold for interindividual
 4    variation, 3-fold for residual toxicodynamic differences associated with animal-to-human
 5    extrapolation,  and 3-fold for database uncertainty) was applied to the POD for the derivation of
 6    the RfC, as described below.
 7    5.1.3.2.1. Interindividual Variation UFH. A factor of 10 was applied to account for variation in
 8    sensitivity within the human population (UFH).  The UF of 10 is commonly considered to be
 9    appropriate in the absence of convincing data to the contrary.  The data from which to determine
10    the potential extent of variation in how humans respond to chronic exposure to methanol are
11    limited, given the complex nature of the developmental endpoint employed and uncertainties
12    surrounding the importance of metabolism to the observed teratogenic effects. Susceptibility to
13    methanol is likely to involve intrinsic and  extrinsic factors. Some factors may include alteration
14    of the body burden of methanol or its metabolites,  sensitization of an individual to methanol
                                                  5-15
       DRAFT—DO NOT CITE OR QUOTE

-------
 1    effects, or augmentation of underlying conditions or changes in processes that share common
 2    features with methanol effects. Additionally, inherent differences in an individual's genetic
 3    make-up, diet, gender, age, or disease state may affect the pharmacokinetics and
 4    pharmacodynamics of methanol, influencing susceptibility intrinsically.  Co-exposure to a
 5    pollutant that alters metabolism or clearance or that adds to background levels of metabolites
 6    may also affect the pharmacokinetics and pharmacodynamics of methanol, influencing
 7    susceptibility extrinsically (see Section 4.9). The determination of the UF for human variation is
 8    supported by several types of information, including information concerning background levels
 9    of methanol in humans, variation in pharmacokinetics revealed through human studies and from
10    PBPK modeling, variation of methanol metabolism in human tissues, and information on
11    physiologic factors (including gender and age), or acquired factors (including diet and
12    environment) that  may affect methanol exposure and toxicity.
13           In using the AUC of methanol in blood as the dose metric for derivation of health
14    benchmarks for methanol, the assumption is made that concentrations of methanol in blood over
15    time are related to its toxicity, either through the actions of the parent or it subsequent
16    metabolism.  However, the formation of methanol's metabolites has been shown in humans to be
17    carried out by enzymes that are inducible, highly variable in activity, polymorphic, and to also be
18    involved in the metabolism of other drugs and environmental  pollutants. Hence, differences in
19    the metabolism of methanol that are specific for target tissue,  gender, age, route of
20    administration, and prior exposure to other environmental chemicals may give a different pattern
21    of methanol toxicity if metabolism is required for that toxicity. Eighty-five percent of Asians
22    carry an atypical phenotype of ADH that may affect their ability to metabolize methanol
23    (Agarwal, 2001; Bosron and Li, 1986; Pietruszko, 1980). Also, polymorphisms in ADH3
24    occurring in the promoter region reduce  the transcriptional activity in vitro nearly twofold,
25    although no studies have reported differences in ADH3 enzyme activity in humans (Hedberg
26    etal.,2001).
27           Although data on the specific potential for increased susceptibility to methanol are
28    lacking, there is information on PK and pharmacodynamic factors suggesting that children may
29    have differential susceptibility to methanol toxicity  (see Section 4.10.1).  Thus, there is
30    uncertainty in children's responses to methanol that should be taken into consideration for
31    derivation of the UF for human variation that is not available  from either measured human data
32    or PBPK modeling analyses. The enzyme primarily responsible for metabolism of methanol in
33    humans, ADH, has been reported to be reduced in activity in newborns. Differences in
34    pharmacokinetics include potentially greater pollutant intake due to greater ventilation rates,
35    activity, and greater intake of liquids in children.  In terms of  differences in susceptibility to
36    methanol due to pharmacodynamic considerations, the substantial anatomical, physiologic, and
37    biochemical changes that occur during infancy, childhood, and puberty suggest that there are
                                                5-16       DRAFT—DO NOT CITE OR QUOTE

-------
 1    developmental periods in which the endocrine, reproductive, immune, audiovisual, nervous, and
 2    other organ systems may be especially sensitive.
 3          There are some limited data from short-term exposure studies in humans and animal
 4    experiments that suggest differential susceptibility to methanol on the basis of gender. Gender
 5    can provide not only different potential targets for methanol toxicity but also differences in
 6    methanol pharmacokinetics and pharmacodynamics. NEDO (1987) reported that in rats exposed
 7    to methanol pre- and postnatally, 6- and 8-week-old male progeny had significantly lower brain
 8    weights at  1,000 ppm, whereas females only showed decreases at 2,000 ppm.  In general, gender-
 9    related differences in distribution and clearance of methanol may result from the greater muscle
10    mass, larger body size, decreased body fat, and increased volumes of distribution in males
11    compared to females.
12    5.1.3.2.2. Animal-to-Human Extrapolation UFA. Afactor of 3 was applied to account for
13    uncertainties in extrapolating from rodents to humans.  Application of a full UF  of 10 would
14    depend on  two areas of uncertainty: toxicokinetic and toxicodynamic uncertainty. In this
15    assessment, the toxicokinetic component is largely  addressed by  the determination of a HEC
16    through the use of PBPK modeling.  Given the chosen dose metric (AUC for methanol blood),
17    uncertainties in the PBPK modeling of methanol are not expected to be greater for one species
18    than another.  The analysis of parameter uncertainty for the PBPK modeling performed for
19    human, mouse, and rat data gave similar results as to how well the model fit the available data.
20    Thus, the human and rodent PBPK model performed similarly using this dose metric for
21    comparisons between species. As discussed in Section 5.3 below, uncertainty does exist
22    regarding the relation of maternal blood levels estimated by the model to fetal and neonatal
23    blood levels that would be obtained under the (gestational, postnatal and lactational) exposure
24    scenario employed in the critical study.  However, at environmentally relevant exposure levels, it
25    is assumed that the ratio of the difference in blood concentrations between a human infant and
26    mother would be similar to and not significantly greater than the difference between a rat dam
27    and its fetus.  Key parameters and factors which determine the ratio of fetal or neonatal human
28    versus mother methanol  blood levels either do not change significantly with age (partition
29    coefficients, relative blood flows) or scale in a way that is common across  species
30    (allometrically). For this reason and because EPA has  confidence in the ability of the PBPK
31    model to accurately predict adult blood levels of methanol, the PK uncertainty is reduced and a
32    value of 1 was applied.  Rodent-to-human pharmacodynamic uncertainty is covered by a factor
33    of 3, as is the practice for deriving RfCs (U.S. EPA, 1994b). Therefore, a factor of 3 is used for
34    interspecies uncertainty.
35    5.1.3.2.3. Database UFD.  A database UF  of 3 was applied to account for deficiencies in the
36    toxicity database. The database for methanol toxicity is quite extensive: there are chronic and

                                               5-17      DRAFT—DO NOT CITE OR QUOTE

-------
 1    developmental toxicity studies in rats, mice, and monkeys, a two-generation reproductive
 2    toxicity study in rats, and neurotoxicity and immunotoxicity studies. However, there is
 3    uncertainty regarding which test species is most relevant to humans. In addition, limitations of
 4    the developmental toxicity database employed in this assessment include gaps in testing and
 5    imperfect study design, reporting, and analyses. Developmental studies were conducted at levels
 6    inducing maternal toxicity, a full developmental neurotoxicity test (DNT) in rodents has not been
 7    performed and is warranted given the critical effect of decreased brain weight, there are no
 8    chronic oral studies in mice, and chronic and developmental studies in monkeys were generally
 9    inadequate for quantification purposes,  for reasons discussed in Section 5.1.1.1. Problems of
10    interpretation of developmental and reproductive studies also arise given the dose spacing
11    between lowest and next highest level.  For these reasons, an UF of 3 was applied to account for
12    deficiencies in the database.
13    5.1.3.2.4.  Extrapolation from Subchronic to Chronic andLOAEL-toNOAEL
14    Extrapolation UFs. AUF was not necessary to account for extrapolation from less than chronic
15    results because developmental toxicity (cervical rib and decreased brain weight) was used as the
16    critical effect. The developmental period is recognized as a susceptible lifestage where exposure
17    during certain time windows is more relevant to the induction of developmental effects than
18    lifetime exposure (U.S. EPA, 1991).
19          A UF for LOAEL-to-NOAEL extrapolation was not applied because BMD analysis was
20    used to determine the POD,  and this factor was addressed as one of the considerations in
21    selecting the BMR. In this case, a BMR of one S.D. from the control mean in the critical effect
22    was selected based on the assumption that it represents a minimum biologically significant
23    change.

      5.1.4. Previous RfC Assessment
24          The health effects data for methanol were assessed for the IRIS database in 1991 and
25    were determined  to be inadequate for derivation of an RfC.

      5.2. ORAL RfD
26          In general, the RfD is an estimate of a daily exposure to the human population (including
27    susceptible subgroups) that is likely to be without an appreciable risk of adverse health effects
28    over a lifetime. It is derived from a POD, generally the statistical lower confidence limit  on the
29    BMDL, with uncertainty/variability factors applied to reflect limitations of the data used.  The
30    RfD is expressed in terms of mg/kg-day of exposure to an agent and is derived by a similar
31    methodology as is the RfC.  Ideally, studies with the greatest duration of exposure and conducted
32    via the oral route of exposure give the most confidence for derivation of an RfD.  For methanol,
33    the oral database  is currently more limited than the inhalation database.  With the  development of
                                               5-18      DRAFT—DO NOT CITE OR QUOTE

-------
 1    PBPK models for methanol, the inhalation database has been used to help bridge data gaps in the
 2    oral database to derive an RfD.

      5.2.1. Choice of Principal Study and Critical Effect-with Rationale and Justification
 3          No studies have been reported in which humans have been exposed subchronically or
 4    chronically to methanol by the oral route of exposure and thus, would be suitable for derivation
 5    of an oral RfD. Data exist regarding effects from oral exposure in experimental animals, but
 6    they are more limited than data from the inhalation route of exposure (see Sections 4.2, 4.3, and
 7    4.4).
 8          Only 2 oral studies of 90-days duration or longer in animals have been reported (Soffritti
 9    et al., 2002a; U.S. EPA, 1986c) for methanol.  EPA (1986c) reported that there were no
10    differences in body weight gain, food consumption, or gross or microscopic evaluations in
11    Sprague-Dawley rats gavaged with 100, 500, or 2,500 mg/kg-day versus control animals. Liver
12    weights in both male and female rats were increased, although not significantly, at the 2,500
13    mg/kg-day dose level, suggesting  a treatment-related response despite the absence of
14    histopathologic lesions in the liver. Brain weights of high-dose group males and females were
15    significantly less than control animals at terminal (90 days) sacrifice. The data were not reported
16    in adequate detail for dose-response modeling and BMD estimation. Based primarily on the
17    qualitative findings presented in this study, the 500 mg/kg-day dose was deemed to be a
18    NOAEL.75
19          The only lifetime oral study available was conducted by Soffiritti et al. (2002a) in
20    Sprague-Dawley rats exposed to 0, 500, 5,000, 20,000 ppm (v/v) methanol,  provided ad libitum
21    in drinking water. Based on default, time-weighted average body weight estimates for Sprague-
22    Dawley rats (U.S. EPA, 1988), average daily doses of 0, 46.6, 466, and 1,872 mg/kg-day for
23    males and 0, 52.9, 529, 2,101 mg/kg-day for females were reported by the study authors. All rats
24    were exposed for up to 104 weeks, and then maintained until natural death.  The authors report
25    no substantial changes in survival nor was there any pattern of compound-related clinical signs
26    of toxicity.  The authors did not report noncancer lesions, and there were no reported compound-
27    related signs of gross pathology or histopathologic lesions indicative of noncancer toxicological
28    effects in response to methanol.
29          Five oral studies investigated the reproductive and developmental effects of methanol in
30    rodents (Aziz et al., 2002; Fu et al., 1996; Sakanashi et al., 1996; Rogers et al., 1993a; Infurna
31    and Weiss, 1986), including three  studies that investigated the influence of FAD diets on the
32    effects of methanol exposures (Aziz et al., 2002; Fu et al., 1996; Sakanashi et al., 1996). Infurna
33    and Weiss (1986) exposed pregnant Long-Evans rats to 2,500 mg/kg-day in drinking water on
34    either GDIS-GDI7 or GD17-GD19. Litter size, pup birth weight, pup postnatal weight gain,
      75 EPA (1986c) did not report details required for a BMD analysis such as standard deviations for mean responses.
                                                5-19      DRAFT—DO NOT CITE OR QUOTE

-------
 1    postnatal mortality, and day of eye opening were not different in treated animals versus controls.
 2    Mean latency for nipple attachment and homing behavior (ability to detect home nesting
 3    material) were different in both methanol treated groups. These differences were significantly
 4    different from controls. Rogers et al. (1993a) exposed pregnant CD-I mice via gavage to 4 g/kg-
 5    day methanol, given in 2 equal daily doses. Incidence of cleft palate and exencephaly was
 6    increased following maternal exposure to methanol. Also, an increase in totally resorbed litters
 7    and a decrease in the number of live fetuses per litter were observed.
 8          Aziz et al. (2002), Fu et al. (1996), and Sakanashi et al. (1996) investigated the role of
 9    folic acid in methanol-induced developmental neurotoxicity.  Like Rogers et al. (1993a), the
10    former 2 studies observed that an oral gavage dose of 4-5 g/kg-day during GD6-GD15 or GD6-
11    GD10 resulted in an increase in cleft palate in mice fed sufficient folic acid diets, as well as an
12    increase in resorptions and a decrease in live fetuses per litter.  Fu et al. (1996) also observed an
13    increase in exencephaly in the FAS group. Both studies found that an approximately  50%
14    reduction in maternal liver folate concentration resulted in an increase in the percentage of litters
15    affected by  cleft palate (as much as threefold) and an increase in the percentage of litters affected
16    by exencephaly (as much as 10-fold).  Aziz et al. (2002) exposed rat dams throughout their
17    lactation period to 0, 1, 2, or 4% v/v methanol via the drinking water, equivalent to
18    approximately 480, 960 and 1,920 mg/kg-day.76 Pups were exposed to methanol via lactation
19    from PND1-PND21. Methanol treatment at 2% and 4% was associated with significant
20    increases in activity (measured as distance traveled in a spontaneous locomotor activity test) in
21    the FAS group (13 and 39%, respectively) and most notably, in the FAD group (33 and 66%,
22    respectively) when compared to their respective controls. At PND45, the CAR in FAD rats
23    exposed to 2% and 4% methanol was  significantly decreased by 48% and 52%, respectively,
24    relative to nonexposed controls. In the FAS group, the CAR was only significantly decreased in
25    the 4% methanol-exposed animals and only by 22% as compared to their respective controls.
      5.2.1.1. Expansion of the Oral Database by Route-to-Route Extrapolation
26          Given the oral  database limitations, including the limited reporting of noncancer findings
27    in the subchronic (U.S. EPA,  1986c) and chronic studies (Soffritti et al., 2002a) of rats and the
28    high-dose levels used in the two rodent developmental studies, EPA has derived an RfD by using
29    relevant inhalation data and route-to-route extrapolation with the aid of the EPA PBPK model
30    (see Sections  3.4 and 5.1).  Several other factors support use of route-to-route extrapolation for
31    methanol.  The limited data for oral administration indicate similar effects as reported via
32    inhalation exposure (e.g., the brain and fetal skeletal system are targets of toxicity). Methanol
33    has been shown to be rapidly and well-absorbed by both the oral and inhalation routes of
      76 Assuming that Wistar rat drinking water consumption is 60 mL/kg-day (Rogers et al., 2002), 1% methanol in
      drinking water would be equivalent to 1% x 0.8 g/mL x 60 mL/kg-day = 0.48 g/kg-day = 480 mg/kg-day.
                                                5-20      DRAFT—DO NOT CITE OR QUOTE

-------
 1    exposure (CERHR, 2004; Kavet and Nauss, 1990). Once absorbed, methanol distributes rapidly
 2    to all organs and tissues according to water content, regardless of route of exposure.
 3          As with the species-to-species extrapolation used in the development of the RfC, the dose
 4    metric used for species-to-species and route-to-route extrapolation of inhalation data to oral data
 5    is the AUC of methanol in blood.  Simulations for human oral methanol exposure were
 6    conducted using the model parameters as previously described for human inhalation exposures,
 7    with human oral kinetic/absorption parameters from Sultatos et al. (2004) (i.e., KAS = 0.2,
 8    KSI = 3.17, and KAI = 3.28). Human oral exposures were assumed to occur during six drinking
 9    episodes during the day, at times  0, 3, 5, 8,  11, and 15 hours from the first ingestion of the day.
10    For example, if first ingestion occurred at 7 am, these would be at 7 am, 10 am, 12 noon, 3 pm, 6
11    pm, and 10 pm.  Each ingestion event was treated as occurring over 3 minutes, during which the
12    corresponding fraction of the daily dose was infused into the  stomach lumen compartment.  The
13    fraction of the total ingested methanol simulated at each of these times was 25%, 10%, 25%,
14    10%, 25%, and 5%, respectively. Six days  of exposure were simulated to allow for any
15    accumulation (visual inspection of plots showed this to be finished by the 2nd or 3rd day), and
16    the results for the last 24 hours were used. Dividing the exposure into more and smaller episodes
17    would decrease the estimated peak concentration but have little effect on AUC. This dose metric
18    was used for dose-response modeling to derive the POD, expressed as a BMDL. The BMDL
19    was then back-calculated using the EPA PBPK model to obtain an equivalent oral drinking water
20    dose in terms of mg/kg-day.

      5.2.2.  RfD Derivation-Including Application ofUFs
      5.2.1.2. Consider a tion oflnhala tion Da ta
21          Inhalation studies considered for derivation of the RfC are used to supplement the oral
22    database using the route-to-route  extrapolation, as previously described.  BMD approaches were
23    applied to the existing inhalation  database, and the EPA PBPK model was used for species-to-
24    species extrapolations. The rationale and approach for determining the RfC is described above
25    (Section 5.1), and the data used to support the derivation of the RfC were extrapolated using the
26    EPA PBPK model to provide an oral equivalent POD.
      5.2.1.3. Selection of Critical Effect(s) from Inhalation Data
27          Methanol-induced effects on the brain in rats (weight decrease) and fetal axial skeletal
28    system in mice (cervical ribs and cleft palate) were consistently observed at lower levels, than
29    other targets, in the oral and inhalation databases.  Analysis of inhalation developmental toxicity
30    studies shows lower BMDLs for decreased  male brain weight in  rats exposed throughout
31    gestation and the FI generation (NEDO, 1987) than BMDLs associated with the fetal axial
32    skeletal system in mice. Therefore, as was noted in Section 5.1.3.1, the BMDL for decreases in

                                               5-21      DRAFT—DO NOT CITE OR QUOTE

-------
 1    brain weight in male rats is chosen to serve as the basis for the route-to-route extrapolation and
 2    cal cul ati on of the RfD.
      5.2.1.4. Selection ofth e POD
 3          The BMDL chosen for the RfC is used to determine the POD for the RfD. This value is
 4    based on a developmental toxicity dataset that includes in utero and postnatal exposures and is
 5    below the range of estimates for other developmental datasets consisting of exposure only
 6    throughout organogenesis.  The neonatal brain is the target organ chosen for derivation of the
 7    RfC. The BMDL for the RfC (AUC of 90.9 hr x mg/L methanol in blood) is converted using the
 8    EPA model to a human equivalent oral exposure of 38.5 mg/kg-day.77

      5.2.2. RfD Derivation-Application of UFs
 9          In an approach consistent with the RfC derivation, UFs are applied to the oral POD of
10    38.5 mg/kg-day to address interspecies extrapolation, intraspecies variability, and database
11    uncertainties for the RfD.  Because the same dataset, endpoint, and PBPK model used to derive
12    the RfC were also used to calculate the oral POD, the total UF of 100 is applied to the BMDL of
13    38.5 mg/kg-day to yield an RfD of 1.12 mg/kg-day for methanol.

14          RfD = 38.5 mg/kg-day + 100 = 0.4 mg/kg-day (rounded to one significant figure)

      5.2.3. Previous RfD Assessment
15          The previous IRIS assessment for methanol included an RfD of 0.5 mg/kg-day that was
16    derived from a  EPA (1986c) subchronic oral study in which Sprague-Dawley rats (30/sex/dose)
17    were gavaged daily with 0, 100, 500, or 2,500 mg/kg-day of methanol. There were no
18    differences between dosed animals and controls in body weight gain, food consumption, gross or
19    microscopic evaluations. Elevated levels of SGPT, serum alkaline phosphatase (SAP), and
20    increased but not statistically  significant liver weights in both male and female rats suggest
21    possible treatment-related effects in rats dosed with 2,500 mg methanol/kg-day, despite the
22    absence of supportive histopathologic lesions in the liver.  Brain weights of both high-dose group
23    males and females were significantly less than those of the control group. Based on these
24    findings, 500 mg/kg-day of methanol was considered a NOAEL in this rat study. Application of
25    a 1,000-fold UF (interspecies extrapolation, susceptible human subpopulations, and subchronic
26    to chronic extrapolation) yielded an RfD of 0.5 mg/kg-day.
      77 The PBPK model used for this HEC estimate is described in Appendix B.  An algebraic equation is provided
      (Equation 2) that describes the relationship between predicted methanol AUC and the HED in mg/kg-day.
                                                5-22       DRAFT—DO NOT CITE OR QUOTE

-------
    5.3. UNCERTANTIES IN THE INHALATION RfC AND ORAL RfD
1          The following is a more extensive discussion of the uncertainties associated with the RfC
2   and RfD for methanol beyond that which is addressed quantitatively in Sections 5.1.2, 5.1.3, and
3   5.2.2. A summary of these uncertainties is presented in Table 5-5.

         Table 5-5. Summary of uncertainties in methanol noncancer risk assessment
Consideration
Choice of endpoint
Choice of dose metric
Choice of model for
BMDL derivation
Choice of animal-to-
human extrapolation
method
Statistical uncertainty
at POD (sampling
variability due to
bioassay size)
Choice of bioassay
Choice of
species/gender
Human population
variability
Potential Impact
Use of other endpoint
could t RfC by up to
~5-fold (see Table 5-4
and Section 5. 3.1)
Alternatives could t
or | RfC/D (e.g., use
of Cmax increased RfC
by -20%)
Use of a linear model
could t RfC by -2.5-
fold (see Table 5-3)
Alternatives could t
or | RfC/D (e.g., use
of standard dosimetry
assumption would t
RfC by -2-fold; see
Section 5. 3. 4)
POD would be -90%
higher if BMD were
used
Alternatives could t
RfC/D
RfC would be t or |
if based on another
species/gender
RfC could | or t if
another value of the
UF was used
Decision
RfC is based on the
most sensitive and
quantifiable endpoint,
decreased brain weight
in male rats exposed
pre- and postnatally
AUC for methanol in
arterial blood
Hill model used
A PBPK model was
used to extrapolate
animal to human
concentrations
A BMDL was used as
the POD
NEDO (1987)
RfC is based on the
most sensitive and
quantifiable endpoint
(I brain weight) in the
most sensitive species
and gender adequately
evaluated (male rats).
10-fold uncertainty
factor applied to derive
the RfC/RfD values
Justification
Chosen endpoint is considered the most
relevant due to its biological significance,
and consistency across a developmental
and a subchronic study in rats and with the
observation of other developmental
neurotoxicities reported in monkeys.
AUC was selected as the most appropriate
dose metric because it incorporates time
(brain weight is sensitive to both the level
and duration of exposure) and better
reflects exposure within a given day.
Hill model gave lowest of a broad range of
BMDL estimates from adequate models
and provides good fit in low dose region.
Use of a PBPK model reduced uncertainty
associated with the animal to human
extrapolation. AUC blood levels of
methanol is an appropriate dose metric and
a peer-reviewed PBPK model that
estimates this metric was verified by EPA
using established (U.S. EPA, 2006a)
methods and procedures
Lower bound is 95% CI of administered
exposure
Alternative bioassays were available, but
the chosen bioassay was adequately
conducted and reported and resulted in the
most sensitive and reliable BMDL for
derivation of the RfC.
Choice of female rats would have resulted
in a higher RfC/D. Effects in mice also
yield higher RfCs. Qualitative evidence
from NEDO (1987) and Burbacher et al.
(2004a, 2004b) suggest that monkeys may
be a more sensitive species, but data are not
as reliable for quantification.
10-fold UF is applied because of limited
data on human variability or potential
susceptible subpopulations, particularly
pregnant mothers and their neonates.
                                             5-23
DRAFT—DO NOT CITE OR QUOTE

-------
      5.3.1. Choice of Endpoint
 1          The impact of endpoint selection on the derivation of the RfC and RfD was discussed in
 2    Sections 5.1.3.1 and 5.2.2.2. Potential RfC values considered ranged from 1.7 to 13.6 mg/m3,
 3    depending on whether neurobehavioral function in male monkeys, brain weight decrease in male
 4    rats, or cervical ribs incidence in mice was chosen as the critical effect for derivation of the POD,
 5    with the former endpoint representing the lower end of the RfC range. The use of other
 6    endpoints, particularly pre-term births identified in the Burbacher et al. (2004a, 2004b, 1999a,
 7    1999b) monkey study, would potentially result in lower reference values,  but significant
 8    uncertainties associated with those studies preclude their use as the basis for an RfC.
 9          Burbacher et al.  (2004a, 2004b, 1999a, 1999b) exposedM fascicularis monkeys to 0,
10    262, 786, and 2,359 mg/m3 methanol 2.5 hours/day, 7 days/week during premating/mating and
11    throughout gestation (approximately 168 days).  They observed a slight but statistically
12    significant gestation period shortening in all exposure groups that was largely due to C-sections
13    performed in the methanol exposure groups "in response to signs of possible difficulty in the
14    maintenance of pregnancy," including vaginal bleeding. As discussed in Sections 4.3.2 and
15    5.1.1.2, there are questions concerning this effect and its relationship to methanol exposure.  An
16    ultrasound was not done to confirm the existence of real fetal or placental problems.
17    Neurobehavioral function was assessed in infants during the first 9 months of life. Two tests out
18    of nine, returned positive results possibly related to methanol exposure. VDR performance was
19    reduced in all treated male infants, and was significantly reduced in the 2,359 mg/m3 group for
20    both sexes and the 786 mg/m3 group for males. However, an overall dose-response trend for this
21    endpoint was only observed in females. As discussed in Section 4.4.2, confidence in this
22    endpoint may have been increased by statistical analyses to adjust for multiple testing (CERHR,
23    2004), but it is a measure of functional deficits in sensorimotor development that is consistent
24    with early developmental CNS effects (brain weight changes discussed above) that have been
25    observed in rats.  The Pagan test of infant intelligence indicated small but not significant deficits
26    of performance (time spent looking a novel faces versus familiar faces) in treated infants.
27    Although these results indicate that prenatal and continuing postnatal exposure to methanol can
28    result in neurotoxicity to the offspring, especially when considered in conjunction with the gross
29    morphological effects noted in NEDO (1987), the use of such statistically borderline results is
30    not warranted in the derivation of the RfC, given the availability of better dose-response data in
31    other species.
32          NEDO (1987) also examined the chronic neurotoxicity of methanol inM fascicularis
33    monkeys exposed to 13.1, 131, or 1,310 mg/m3 for up to 29  months. Multiple effects were noted
34    at 131 mg/ m3, including slight myocardial effects (negative changes in the T wave on an EKG),
35    degeneration of the inside nucleus of the thalamus, and abnormal pathology within the cerebral
36    white tissue in the brain. The results support the identification of 13.1 mg/m3 as the NOAEL for
                                                5-24      DRAFT—DO NOT CITE OR QUOTE

-------
 1    neurotoxic effects in monkeys exposed chronically to inhaled methanol.  However, as discussed
 2    in Section 4.2.2.3, there exists significant uncertainty in the interpretation of these results and
 3    their utility in deriving an RfC for methanol.  These uncertainties include lack of appropriate
 4    control group data, limited nature of the reporting of the neurotoxic effects observed, and use of
 5    wild-caught monkeys in the study.  Thus, while the NEDO (1987) study suggests that monkeys
 6    may be a more sensitive species to the neurotoxic effects of chronic methanol exposure than
 7    rodents, the substantial deficits in the reporting of data preclude the quantification of data from
 8    this study for the derivation of an RfC.
 9           The increased incidence of cervical ribs was identified as a biologically significant,
10    potential co-critical effect based on the findings of Rogers et al. (1993a). Mice were exposed to
11    1,000, 2,000, or 5,000 ppm, and incidence of cervical ribs was statistically increased at
12    2,000 ppm.  However, given that the reference values for the increased incidence of cervical ribs
13    are estimated to be approximately five times higher than the reference values calculated using
14    decreases in brain weight in male rats (NEDO,  1987), decreased brain weight was chosen as the
15    basis for the derivation of the RfC.

      5.3.2. Choice of Dose Metric
16           A recent review of the reproductive and developmental toxicity of methanol by a panel of
17    experts concluded that methanol, not its metabolite formate, is likely to be the proximate
18    teratogen and that blood methanol level is a useful biomarker of exposure (CERHR, 2002;
19    Dormanetal., 1995). The CERHR Expert Panel based their assessment of potential methanol
20    toxicity on an assessment of circulating blood levels (CERHR,  2002).  In contrast to the
21    conclusions of the NTP-CERHR panel, in vitro data from Harris et al. (2004, 2003) suggest that
22    the etiologically important substance for embryo dysmorphogenesis and embryolethality was
23    likely to be formaldehyde rather than the parent compound or formate. Although there remains
24    uncertainty surrounding the identification of the proximate teratogen of importance (methanol,
25    formaldehyde, or formate), the dose metric chosen for derivation of an RfC was based on blood
26    methanol levels.  This decision was primarily based on  evidence that the toxic moiety is not
27    likely to be the formate metabolite of methanol (CERHR, 2004), and evidence that levels of the
28    formaldehyde metabolite following methanol maternal  and/or neonate exposure would be lower
29    in the fetus and neonate than in adults. While recent in vitro evidence indicates that
30    formaldehyde is more embryotoxic than methanol and formate, the high reactivity of
31    formaldehyde would limit its unbound and unaltered transport as free formaldehyde from
32    maternal to fetal blood (Thrasher and Kilburn, 2001) (see discussion in Section 3.3). Thus, even
33    if formaldehyde is ultimately identified as the proximate teratogen, methanol would likely play a
34    prominent role, at least in terms of transport to the target tissue. Further discussions of methanol
35    metabolism, dose metric selection, and MOAissues are in Sections 3.3, 4.6, 4.8 and 4.9.2.
                                               5-25       DRAFT—DO NOT CITE OR QUOTE

-------
 1          There exists some concern in using the FI generation NEDO (1987) rat study as the basis
 2    from which to derive the RfC. This concern mainly arises from issues related to the low
 3    confidence that the PBPK model is accurately predicting dose metrics for neonates exposed
 4    through multiple and simultaneous routes.  The PBPK model was structured to predict internal
 5    dose metrics for adult NP animals and was optimized using adult metabolic and physiological
 6    parameters. Young animals have very different metabolic and physiological profiles than adults
 7    (enzyme activities, respiration rates, etc.).  This fact, coupled with multiple routes of exposure,
 8    make it likely that the PBPK did not accurately predict the internal dose metrics for the offspring.
 9    Stern et al. (1996) reported that when rat pups and dams were exposed together during lactation
10    to 4,500 ppm methanol in air, methanol blood levels in pups from GD6-PND21 were
11    approximately 2.25 times greater than those of dams. This discrepancy persisted until PND48,
12    when postnatal exposure continued to PND52.  It is logical to assume that similar differences in
13    blood methanol levels would  also be observed in the NEDO (1987) FI study, as the exposure
14    scenario is similar to that of Stern et al. (1996).  Differences between pup and dam blood
15    methanol  levels might be expected to be slightly greater than twofold in the NEDO (1987) FI
16    study as the exposure was continuous  (versus 6 hours/day in the Stern et al. [1996] paper) and
17    lasted for a longer duration (-64 days  versus 37). Under a similar scenario, human newborns
18    may experience higher blood  levels than their mothers as a result of breast feeding. As has been
19    discussed in Chapter 3, children have a limited capacity to metabolize methanol via ADH;
20    however, there is some evidence that human infants are able to efficiently eliminate methanol at
21    high-exposure levels, possibly via CAT (Tran et al., 2007). At environmentally relevant
22    exposure levels, it is assumed that the  ratio of the difference in blood concentrations between
23    infant and mother would not be significantly greater than the twofold difference that  has been
24    observed in rats.78 For this reason and because EPA has confidence in the ability of the PBPK
25    model to accurately predict adult blood levels of methanol, the maternal blood methanol levels
26    for the estimation of HECs from the NEDO (1987) study were used as the dose metric.

      5.3.3. Choice of Model for BMDL Derivations
27          The Hill model adequately fit the dataset (goodness-of-fit^-value = 0.84). Data points
28    were well predicted near the BMD (scaled residual = 0.09) (see Figure 5-1).  There is a 2.5-fold
29    range of BMDL estimates from adequately fitting models, indicating considerable model
30    dependence. The BMDL from the Hill model was selected, in accordance with EPA BMD
31    Technical Guidance  (EPA, 2000b), because it results in the lowest  BMDL from among a broad
32    range of BMDLs and provides a superior fit in the low dose region nearest the BMD.
      78 Key parameters and factors which determine the ratio of fetal or neonatal human versus mother methanol blood
      levels either do not change significantly with age (partition coefficients, relative blood flows) or scale in a way that
      is common across species (allometrically).
                                                5-26      DRAFT—DO NOT CITE OR QUOTE

-------
      5.3.4. Choice of Animal-to-Human Extrapolation Method
 1           APBPK model developed by the EPA, adapted from Ward et al. (1997), was used to
 2    extrapolate animal-to-human concentrations. An AUC blood level of methanol (90.9 hr x mg/L)
 3    associated with a one S.D. change from the control mean for brain weights in rats was estimated
 4    using the rat PBPK model. Then the human PBPK model was used to convert back to a human
 5    equivalent exposure concentration or a BMCLHEc/isD of 182 mg/m3. If no PBPK models were
 6    available, a BMCLHEc/isD of 424 mg/m3 would have been derived by adjusting the 556.5 mg/m3
 7    BMCLiso for external exposure concentration for duration and the animal-to-human standard
 8    adjustment factor for systemic effects (the ratio of animal and human blood:air partition
 9    coefficients). This value is approximately 2-fold higher than the value derived using the PBPK
10    model.  However, as discussed above, use of PBPK-estimated maternal blood methanol levels
11    for the estimation of HECs allows for the use of data-derived extrapolations rather than standard
12    methods for extrapolations from external exposure levels.
13          As discussed in Section 3.4, the PBPK models do not describe or account for background
14    levels of methanol, formaldehyde or formate, and background levels were subtracted from the
15    reported data before use in model fitting or validation (if not already subtracted by study
16    authors), as described below. This approach was taken because the relationship between
17    background doses and background responses is not known, because the primary purpose of this
18    assessment is for the determination of noncancer and cancer risk associated with increases in the
19    levels of methanol or its metabolites (e.g., formate, formaldehyde) over background, and because
20    the subtraction of background levels is not expected to have a significant impact on PBPK model
21    parameter estimates as background levels of methanol and its metabolites are low relative to
22    exposure levels used in methanol bioassays.

      5.3.5. Route-to-Route Extrapolation
23          To estimate an oral dose POD for decrease in brain weight in rats, a route-to-route
24    extrapolation was performed on the inhalation exposure POD used to derive the RfC.  One way
25    to characterize the uncertainty associated with this approach is to compare risk levels (BMDL
26    values) using the dose metric, AUC methanol, for developmental decreases in brain weight
27    derived from 1) an existing oral subchronic study and 2) from a model estimating this metric
28    from an existing inhalation subchronic study. There are currently no oral developmental studies
29    investigating decreases in brain weight available to compare to the risk values estimated using
30    the second procedure. However, the fact that the oral BMDL of 38.5 mg/kg-day estimated in this
31    assessment from the NEDO (1987) inhalation study of neonate rats via a PBPK model is lower
32    than the NOAEL of 500 mg/kg-day identified in EPA (1986c) methanol  study of adult rats is
33    consistent with other studies which suggest that fetal/neonatal organisms are a sensitive
34    subpopulation.
                                              5-27       DRAFT—DO NOT CITE OR QUOTE

-------
      5.3.6. Statistical Uncertainty at the POD
 1          There is uncertainty in the selection of the BMR level. For decreased brain weight in
 2    rats, no established standard exists, so a BMR of one S.D. change from the control mean was
 3    used. Parameter uncertainty can be assessed through CIs. Each description of parameter
 4    uncertainty assumes that the underlying model and associated assumptions are valid. For the
 5    Hill model applied to the data for decreased brain weight in rats, there is a degree of uncertainty
 6    at the one S.D. level (the POD for derivation of the RfC), with the 95% one-sided lower
 7    confidence limit (BMDL) being -50% below the maximum likelihood estimate of the BMD.

      5.3.7.  Choice of Bioassay
 8          The NEDO (1987) study was used for development of the RfC and RfD because it
 9    resulted in the lowest BMDL. It was also a well-designed study, conducted in a relevant species
10    with an adequate number of animals per dose group, and with examination of appropriate
11    developmental toxicological endpoints. Developmental (Burbacher et al., 2004a, 2004b, 1999a,
12    1999b) and chronic studies (NEDO, 1987) of methanol have been performed in monkeys.  As
13    discussed above in Section 5.3.1 and other sections of this assessment, while the monkey may be
14    a sensitive species for use in the determination of human risk, reporting deficits and study
15    uncertainties preclude their use in the derivation of an RfC.

      5.3.8. Choice of Species/Gender
16          The RfC and RfD were based on decreased brain weight at 6 weeks postbirth in male rats
17    (the gender most sensitive to this effect) (NEDO, 1987).  This decrease in brain weight also
18    occurs in female rats; however, if the decreased brain weight  in female rats had been used, higher
19    RfC and RfD values would have been derived (approximately 66% higher than the male derived
20    values).

      5.3.9. Human Population Variability
21          The extent of interindividual variation of methanol metabolism in humans has not been
22    well characterized. As discussed in Section 4.9, there are a number of issues that may lead to
23    sensitive human subpopulations.  Potentially sensitive  subpopulations would include individuals
24    with polymorphisms in the enzymes involved in the  metabolism of methanol and individuals
25    with significant folate deficiencies. Sensitive lifestages would include children and neonates, as
26    they have increased respiration rates compared to adults, which may increase their methanol
27    blood levels compared to adults.  Also, children have been shown to have decreased ADH
28    activity relative to adults, thus decreasing their ability to metabolize and eliminate methanol. As
29    demonstrated by these examples, there exists considerable uncertainty pertaining to human

                                               5-28      DRAFT—DO NOT CITE OR QUOTE

-------
 1    population variability in methanol metabolism, which provides justification for the 10-fold
 2    intraspecies UF used to derive the RfC and RfD.

      5.4.5.4. CANCER ASSESSMENT

      5.4.1. Oral Exposure

      5.4.1.1. Choice of Study/Data—with Rationale and Justification
 3          No human data exist that would allow for quantification of the cancer risk of chronic
 4    methanol exposure. Table 4-34  summarizes the available experimental animal oral exposure
 5    studies of methanol.  The Soffritti et al. (2002a) and Apaja (1980) oral studies report effects that
 6    show a statistically significant increase in incidence of cancer endpoints in the treated groups
 7    versus the control group (pair-wise comparison). As detailed in Section 4.2.1.3, Soffritti et al.
 8    (2002a) exposed Sprague-Dawley rats via drinking water to 500-20,000 ppm methanol for
 9    104 weeks. Exposure ended at 104 weeks, but the animals were not euthanized and were
10    followed until their natural death.  Increased lymphoma responses in multiple organs of male and
11    female rats were the only carcinogenic effects reported in the Soffritti et al. (2002a) methanol
12    drinking water study that are considered dose related and quantifiable.  Hepatocellular
13    carcinomas observed in male rats are considered potentially dose related (relative to historical
14    controls) but are not quantifiable due to the lack of a statistically significant dose-response trend.
15    Significant increases reported for head and ear duct carcinomas in male rats were not used
16    because NTP pathologists interpreted a majority of these ear duct responses as being
17    hyperplastic, not carcinogenic, in nature (EFSA, 2006; Hailey, 2004).  Apaja (1980) observed
18    significant increases in malignant lymphomas relative to untreated, historical controls in Swiss
19    mice exposed to methanol in drinking water for life. Due to the lack of a concurrent control, the
20    Apaja (1980) study was not considered adequate for derivation of an oral slope factor, but a
21    quantitative analysis  of the dose-response data from this study is included in Appendix E for
22    comparison purposes.
      5.4.1.2. Dose-Response Data
23          The tumor incidence data selected for modeling were the lympho-immunoblastic
24    lymphomas and the combined lympho-immunoblastic, lymphoblastic and lymphocytic
25    lymphomas in both male and female rats of the Soffritti et al (2002a) study. These lymphomas
26    were combined at the recommendation of NTP pathologists  due to their similar histological
27    origin. The incidence of histiocytic sarcomas and myeloid leukemias was not significantly
28    increased in either sex, and the data for these tumors was not combined with the lymphoblastic
29    lymphomas because they are of a different cell line and the combination is not typically
30    evaluated either for statistical significance or dose-response modeling  (Hailey, 2004; McConnell,
31    et al., 1986). Table 5-6 gives the lymphoma incidence data from the study which differs slightly


                                               5-29      DRAFT—DO NOT CITE OR QUOTE

-------
 1    from the data reported in Soffritti et al. (2002a) in the incidence of lympho-immunoblastic
 2    lymphomas in the male 5,000 ppm group.79

          Table 5-6. Incidence data for lymphoma, lympho-immunoblastic, and all lymphomas
          in male and female Sprague-Dawley rats
Dose (ppm)
Dose
(mg/kg-day)
Number of animals
examined
Lymphoma lympho-
immunoblastic
All lymphomas
combined
Female rats
0
500
5,000
20,000
0
66.0
624.1
2,177
100
100
100
100
9
17
19a
21a
9
19a
20a
22b
Male rats
0
500
5,000
20,000
0
53.2
524
1,780
100
100
100
99
16
24
28a
37b
17
27
29a
38b
       Statistically significant by Fisher's Exact test: *p < 0.05, > < 0.01
       Source:  Soffritti et al. (2002a) and ERF web portal (http://www.ramazzini.it/fondazione/foundation.aspX

      5.4.1.3. Dose A djustm en ts an d Ext rap ola tion Meth od
 3          As with the extrapolations used in the development of the RfC and RfD, the PBPK model
 4    was used for species-to-species extrapolation of the doses to be used in the cancer dose-response
 5    analysis.  Three dose metrics were considered for use in the dose-response analysis: total
 6    metabolized methanol; maximum blood concentration of the parent (Cmax); and area under the
 7    blood concentration time curve (AUC) for the parent. Internal dose estimates (above
 8    background) corresponding to the administered doses from the animal bioassay were determined
 9    for each of these metrics with the PBPK model  (see Appendix E, Table E-5). To help inform the
10    selection of the most appropriate dose metric, dose-response analyses were performed using
11    these PBPK model results to assess which dose metric best corresponded to the observed
12    incidence data in Table 5-6 (see Appendix E, Tables E-6 and E-7).  Figures 5-3, 5-4, and 5-5
13    show the fit of the multistage model to the all lymphoma incidence data for female and males,
14    using each dose metric as the dose input.
      79 EPA obtained detailed, individual animal data via an interagency agreement with NIEHS which supported the
      development of reports made available through the Ramazzini Foundation (ERF) web portal
      (http://www.ramazzini.it/fondazione/foundation.asp).  This allowed EPA to combine lymphomas of similar
      histopathological origin and confirm the tumor incidences reported in the Soffriti et al. (2000a) paper.
                                                5-30       DRAFT—DO NOT CITE OR QUOTE

-------
                         Multistage Cancer Model
                     "Multistage Cat
                    Female rats (p = 0.18)
             0    20    40   60
                                   100   120   140
    22:46 07/23 2009
                                                                         Multistage Cancer Model
                                                    <     0.3
               Male rats (p = 0.2)
                                                                     50       100       150       200
                                                     22:51 07/23 2009
Figure 5-3. All lymphomas versus methanol metabolized (mg/day) for female and male
rats
                      Multistage Cancer Model
                 Female rats (p = 0.077)
  22:3507/232009
                       2000    3000
                          dose
                                     4000     5000
0.45


 0.4


0.35


 0.3


0.25


 0.2


0.15
                                                                      Multistage Cancer Model
                                                                   Multistage Cancer
Male rats (p = 0.15)
                                                           0   500  1000 1500  2000  2500  3000  3500 4000  4500
Figure 5-4. All lymphomas versus Cmax (mg/L) for female and male rats.
                     Multistage Cancer Model                                      Multistage Cancer Model
                  Multistage Cancer
                Female rats (p = 0.075)
         0    20000    40000    60000   80000   100000
 22:41 07/232009
                                                  22:44 07/23 2009
            Male rats (p = 0.15)
                                                                20000    40000    60000    80000    100000
                                                                          dose
Figure 5-5. All lymphomas versus AUC (hr x mg/L) for male and female rats
                                              5-31        DRAFT—DO NOT CITE OR QUOTE

-------
 1          The dose-response modeling suggests that total metabolized methanol is a better dose
 2    metric than the parent compound metrics as indicated by improved model fit to responses
 3    reported by Soffritti et al. (2002a) for both lympho-immunoblastic lymphoma (see Appendix E,
 4    Table E-7) and all lymphoma (see Figures  5-3 to 5-5 and Appendix E, Table E-7), and also for
 5    malignant lymphoma responses reported by Apaja (1989) (see Appendix E, Table E-17).  Chi-
 6    square p values for the total metabolite dose metric ranged from 0.18 to 0.55 and were
 7    consistently higher than for the other dose  metrics. This could be an indication of the importance
 8    of metabolite formation, which is likely to be more rapid at low doses, to the carcinogenic
 9    response.  The total metabolized methanol dose metric was selected as the dose metric for use in
10    the dose-response assessment to derive the POD because it provided the best fit to the response
11    data. The estimated BMDL for the total methanol metabolized dose metric was then back-
12    calculated using the EPAPBPK model to obtain a human equivalent oral drinking water dose in
13    terms of mg/kg-day (see Appendix E, Table E-8).
14          Multistage and multistage Weibull  time-to-tumor models were applied to the lymphoma
15    data obtained from ERF for the Sofffritti et al. (2002a) drinking water study and considered for
16    determining the POD to be used in the derivation of the oral cancer slope factor (see Appendix E,
17    Table 3-8). Appendix E gives the details and justification for the various approaches used. As
18    described in Appendix E, time-to-tumor modeling and multistage quantal modeling gave similar
19    results, and the tumor responses modeled did not exhibit significant time dependence on dose.
20    The EPA multistage cancer model fit the response data adequately and was used to derive the
21    oral cancer slope factor (CSF) (see Appendix E, Tables E-7, E-8, and Figure E-10).
22          BMDs and BMDLs were estimated for the combined lymphomas in male and female rats.
23    The BMR selected was the standard value  of 10% extra risk recommended for dichotomous
24    models (U.S. EPA, 2000b).80 The 95% one-sided lower confidence limit defined the BMDL.
25    The dose terms in the fitting were set equal to the estimated total metabolized doses derived
26    using the PBPK model for methanol for each of the administered doses in the bioassay.
27          Application of the multistage model to the incidence data for all lymphomas in male  rats
28    (Table 5-6) resulted in the BMD and BMDLio values presented in Table 5-7. The results for the
29    male rat were used in the derivation of the CSF because the female data for this endpoint yielded
30    slightly higher values (see Appendix E, Tables E-7 and E-8). Assuming that metabolized
31    methanol distributes in the body according to body weight to the % power, the male rat mg-day
32    BMDLio was converted to a human mg-day BMDLio.81 The human PBPK model (Appendix B)
33    was then used to convert this human mg-day value for total methanol metabolized back to a
      80 The use of lower BMR values was determined not to have a significant impact on the CSF derivation.
      81 Male rat mg/day was converted to human mg/day by multiplying by (BWhuman)yV(BWrat)y4 =(70 kg)yY(0.33 kg f'=
      55.6
                                               5-32       DRAFT—DO NOT CITE OR QUOTE

-------
 1   human equivalent methanol oral dose HED(BMDLio) of 51.5 mg/kg-day for lymphomas in the
 2   male rat (see Appendix E, Table E-8).82

          Table 5-7. BMD results and oral CSF using all lymphoma in male rats
Amount metabolized (mg/d)
BMD10
(mg/d)
101.7
Rat BMDL10
(mg/d)
63.9
Estimated human
BMDL10 (mg/d)
3,553
Human equivalent
BMDL10
(mg/kg-day)
51.5
Oral CSF
(mg/kg-day) 1
1.9E-03
 4
 5
 6
 7
 9
10
11

12
13
14
15
16
17
18
19
Source: Soffritti et al. (2002a).

       In the case of methanol, there is no information to inform the MOA for carcinogenicity.
As recommended in the Guidelines for Carcinogen Risk Assessment (U.S. EPA, 2005a), "when
the weight of evidence evaluation of all available data is insufficient to establish the MOA for a
tumor site and when scientifically plausible based on the available data, linear extrapolation is
used as a default approach." Accordingly, for the derivation of a quantitative estimate of cancer
risk for ingested methanol, a linear extrapolation was performed to determine the CSF.
5.4.1.4. Oral Slope Fa ctor
       The oral slope factor was derived based on a linear extrapolation from this POD
(BMDLio/HED of 55.4 mg/kg-day for lymphomas in the male rat) to the estimated background
response level:

               O.I/HED(BMDLio) = 0.1/51.5 mg/kg-day = 2E-03 (mg/kg-day)'1
                            (rounded to one significant figure)

5.4.2. Inhalation Exposure
5.4.2.1. Choice of Study/Data-with Rationale and Justification
       No human data exist that would allow for quantification of the cancer risk associated with
chronic methanol exposure. Table 4-35 summarizes the available experimental animal inhalation
exposure studies of methanol. The NEDO (1987,  1985/2008b) 24-month rat study is the only
inhalation bioassay available that reports an increase in incidence of any cancer endpoints (see
Section 4.2.2.3). This NEDO (1987, 1985/2008b) study was of high quality and was based on
standard OECD guidelines (OECD, 2007).  F344 rats were exposed for 104 weeks to air
concentrations of 0, 10, 100 and 1,000 ppm methanol. Rats were  sacrificed and necropsied at the
     82 The following algebraic equation is provided in Appendix B (Equation 4) to describe the relationship between
     predicted human mg-day methanol cleared and the human equivalent oral dose (HED) in mg/kg-day:
              HED = (3.3*3,553 mg/d)/(19,282.9 - 3, 553 mg/d) + (0.014287*3, 553 mg-d) = 51.5 mg/kg-d
                                               5-33
                                                    DRAFT—DO NOT CITE OR QUOTE

-------
 1    end of the 104-week exposure period. The NEDO (1987, 1985/2008b) study reports increased
 2    pulmonary adenomas/adenocarcinomas and pheochromocytomas in high-dose (1,000 ppm) male
 3    and female rats, respectively. The combined incidence of pulmonary adenomas and
 4    adenocarcinomas was significantly increased in the high-dose males (see Tables 4-4 and 5-8),
 5    and both tumor types were considerably elevated at the high-dose over historical control
 6    incidences within their respective sex and strain (see discussion in Section 4.2.2.3). As shown in
 7    Table 4-5, the severity and combined incidence of potential precursor effects in the alveolar
 8    epithelium of male rat lungs (epithethial swelling, adenomatosis,  pulmonary adenoma, and
 9    pulmonary adenocarcinoma) and the adrenal glands of female rats (hyperplasia and
10    pheochromocytoma) were increased in the higher exposure groups compared with the controls
11    and lower exposure groups. While the incidence of male rat pulmonary adenomas was also high
12    in the lowest (10 ppm) exposure group, the appearance of a rare adenocarcinoma in the high-
13    dose group is suggestive of a progressive effect associated with methanol exposure. While the
14    increased pheochromocytoma response in female rats is not statistically increased over controls,
15    it is considered to be potentially treatment related because this is  a historically rare tumor type
16    for female F344 rats (NTP, 2007, 1999; Haseman et al.,  1998),83 and when viewed in  conjunction
17    with the increased medullary hyperplasia observed in the mid-exposure (100 ppm) group
18    females, it is suggestive of a proliferative change with increasing methanol exposure.
      5.4.2.2. Dose-Response Data
19          The tumor incidence data selected for modeling are the NEDO (1987, 1985/2008b)
20    reported incidences of adenoma/adenocarcinoma in male rats and pheochromocytoma in female
21    rats. These data are presented in Table 5-8.

          Table 5-8. Incidence data for tumor responses in male and female F344 rats
Dose (ppm)


0
10
100
1000

0
10
100
1000
Number of animals affected/number examined
Pheochromocytoma
Pulmonary adenoma/adenocarcinoma
Female rats
2/50
3/51
2/49
7/51a'b
2/52
0/19
0/20
0/52
Male rats
7/52
2/16
2/10
4/51
1/52
5/50
2/52
7/52b,c
      83
        Haseman et al., (1998) report rates for spontaneous pheochromocytomas in 2-year NTP bioassays of 5.7%
      (benign) and 0.3% (malignant) in male F344 rats and 0.3% (benign) and 0.1% (malignant) in female (n=1517) F344
      rats.
                                               5-34      DRAFT—DO NOT CITE OR QUOTE

-------
     I  Dose (ppm) |	Number of animals affected/number examined	
     ap < 0.05 over NTP historical controls for total (benign, complex and malignant) pheochromocytomas using the
     Fisher's Exact test
     bp < 0.05 for Cochrane-Armitage test of overall dose-response trend.
     cp < 0.05 over concurrent controls using the Fisher's Exact test.

     Source: NEDO (1987, 1985/2008b).

     5.4.2.3. Dose A djustm en ts an d Ext rap ola tion Met A od
 1          As with the extrapolations used in the development of the RfC and RfD, the PBPK model
 2   was used for  species-to-species extrapolation of the doses to be used in the cancer dose-response
 3   analysis.  Three  dose metrics were considered for use in the dose-response analysis: total
 4   metabolized methanol, maximum blood concentration of the parent (Cmax), and area under the
 5   blood concentration time curve (AUC) for the parent. Each of the dose metrics corresponding to
 6   the administered dose from the animal bioassay was determined with the PBPK model (see
 7   Appendix E,  Table E-10). To help inform the selection of the most appropriate dose metric,
 8   dose-response analyses were performed using these PBPK model results to assess which dose
 9   metric best corresponded to the observed incidence data in Table 5-8 (see Table E-l 1). All of
10   the dose metrics resulted in similar fit to the incidence data for both endpoints, with the total
11   metabolites metric providing a slightly improved fit to the female pheochromocytoma response
12   data.
13          Unlike the oral data discussed in Section 5.4.1.2, dose-response modeling of the
14   inhalation data from NEDO (1987, 1985/2008b) does not suggest the use of any one dose metric
15   over the other. However, since the pheochromocytoma response likely involves systemically
16   distributed, metabolized methanol, and to be consistent  with the oral CSF, analysis the total
17   methanol metabolized dose metric is selected as the dose metric for use in the dose-response
18   assessment to derive the inhalation POD. The estimated BMDL for the methanol metabolized
19   dose metric was then back-calculated using the EPA PBPK model to obtain a human equivalent
20   air exposure concentration in terms of mg/m3 (see Table E-l3).
21          The EPA multistage model was applied to the data in Table 5-8 obtained from the NEDO
22   (1987, 1985/2008b) inhalation study and considered for determining POD to be used in the
23   derivation of the inhalation cancer unit risk (Table E-l 1). Appendix E gives the details and
24   justification for the various approaches used.  As described in Appendix E, time-to-tumor and
25   quantal modeling gave similar results, and the tumor responses modeled did not exhibit
26   significant time  dependence on dose. The EPA multistage cancer model fit the response data
27   adequately and was used to derive the IUR (Tables E-l 1, E-12,  and  Figure E-13).
28          BMDs and BMDLs were estimated for tumor responses in male and female rats as shown
29   in Table 5-8.  The BMR selected was the standard value of 10% extra risk recommended for

                                               5-35      DRAFT—DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
                                     84
quantal models (U.S. EPA, 2000b).  The 95% one-sided lower confidence limit defined the
BMDL. The dose terms in the fitting were set equal to the estimated total metabolized doses
derived using the PBPK model for methanol for each of the administered doses in the bioassay.
       Application of the multistage model to the incidence data for pheochromocytomas in
female rats (Table 5-8) resulted in the BMD and BMDLio values presented in Table 5-9.  The
results for the female rat were used because the female data for pheochromocytoma yielded
slightly lower BMDL values (Tables E-ll and E-13). Assuming that metabolized methanol
distributes in the body according to body weight to the 3/4 power, the female rat mg-day BMDLio
was converted to a human mg-day BMDLio.85 The human PBPK model (Appendix B) was then
used to convert this human mg-day value for total methanol metabolized back to a human
equivalent methanol inhalation concentration, HEC(BMCLio), of 81.9 mg/m3 or 81,900 ug/m3
for pheochromocytomas in the female rat (see Appendix E, Table  E-13).86

     Table 5-9. BMD results and IUR using pheochromocytoma in female rats
Amount metabolized (mg/day)
BMD10
(mg/day)
29.5
BMDL10
(mg/day)
14.6
Estimated human
BMDL10 (mg/day)
971
Human equivalent
BMCL10
(mg/m3)
81.9
IUR
(jig/m3)-1
1.2E-06
     Source: NEDO (1987, 1985/2008b).

     5.4.2.4. IUR
13          The IUR in terms of (ug/m3)-1 was then derived based on a linear extrapolation from this
14   POD to the estimated background response level:
15
16
17
18
                    0.1/HEC(BMCLio) = 0.17(81,900 ug/m3) = 1E-06 (ug/m3)-1
                             (rounded to one significant figure)

5.4.3. Uncertainties in Cancer Risk Assessment
       The following is a discussion of the uncertainties associated with the cancer potency
estimate for methanol beyond that which can be addressed with the quantitative approach
applied. A summary of these uncertainties is presented in Table 5-10.
     84 The use of lower BMR values was determined not to have a significant impact on the IUR derivation.
     85 Female rat mg/day is converted to human mg/day by multiplying by (BWhuman)/4/(BWrat)y4 = (70 kg)y7(0.26 kg f' =
     66.5
     86 The following algebraic equation is provided in Appendix B (Equation 3) to describe the relationship between
     predicted human mg-day methanol cleared and the human equivalent inhalation concentration (HEC) in mg/m3:
               HED = (14.69 x 971 mg/d)/(19282.9 - 971 mg/d) + (0.063554 x 971 mg/d) = 81.9 mg/ m3

                                               5-36       DRAFT—DO NOT CITE OR QUOTE

-------
         Table 5-10. Summary of uncertainty in the methanol cancer risk assessment
Consideration
Quality of the
studies relied
upon for the
determination of
the PODs
Interpretation of
results from
study relied upon
for the
determination of
the POD
Consistency of
results across
chronic studies
Choice of
endpoint for
POD derivation
Choice of
species/gender
Choice of model
for POD
derivation
Choice of
animal-to-human
extrapolation
method
Potential impact
Key chronic studies
not always well
reported; could lead
to t or I of risks
Differences in tumor
classification can
lead to over or
underestimate of risk
If effects not relevant
to humans, risk is
overestimated.
Route-to-route
extrapolation from
Soffriti et al. (2002a)
study would t
inhalation POD by
about 4-fold
CSF and IUR would
be I if based on
another gender
Use of other models
could t or | POD,
but not significantly
Traditional method
could t the
HEC(BMCL10)
estimate by 4-fold.
Decision
Utilize re-analyses of
the Soffritti et al.
(2002a) and NEDO
(1985/2008a, 2008b)
chronic studies
Derive POD based on
incidence of combined
lymphomas as
suggested by NTP
pathologists; Assume
proper classification of
lung and adrenal
tumors by NEDO
Derive PODs based on
Soffritti et al. (2002a)
and NEDO
(1985/2008b).
Derive oral CSF from
lymphoma data and
inhalation CSF from
adrenal effects.
CSF and IUR are
based on the most
sensitive and reliably
quantifiable
species/gender
Derive cancer potency
factor based on
multistage model.
A PBPK model was
used to extrapolate
animal-to-human
concentrations.
Justification
Consideration of all available information resulted
in the determination that the Soffritti et al. (2002a)
and NEDO (1985/2008a, 2008b) chronic studies
are adequate (see discussion of individual studies
in Sections 4.2.1.3 and 4.2.2.3 and summaries in
Sections 4.9. land 4.9.2).
Both NTP and EFSA recommend that only
lymphomas of the same cellular origin be
combined for dose-response analysis. With
respect to lung and adrenal tumors, examination of
concurrent alveolar and adrenal noncancer
hyperplastic endpoints supports a proliferative
change in these organ systems consistent with the
appearance of carcinogenic responses.
Though tissue concordance across species, strains
and routes of exposure is not assumed, lymphomas
have been observed in more than one species by
oral route. Also there is evidence that the observed
lymphomas are relevant to humans (see
discussions in Section 4.9. concerning human
studies of methanol metabolite formaldehyde).
Oral POD was based on the only tumor type from
Soffritti et al. (2002a) drinking water study
significantly increased (all lymphomas); Inhalation
POD based on most sensitive tumor response from
NEDO (1985/2008b) study, increased
pheochromocytoma in female rats.
Choice of female rat lymphoma and male rat
adenoma/adenocarcinoma would have resulted in
lower CSF and IUR values, respectively. Use of
the Apaja (1989) mouse data would have resulted
in a higher, but less reliable CSF due to study
problems, including a lack of concurrent controls
Use of the multistage model is consistent with
EPA guidance (U.S. EPA, 2005a). The multistage
model provides adequate fit to the data, which is
not improved by a time-to-tumor modeling.
Use of a PBPK model reduces uncertainty
associated with the animal to human extrapolation.
Total metabolites normalized by body weight is an
appropriate dose metric and a verified PBPK
model exists that estimates this metric.
    5.4.3.1. Quality of Studies that are the Basis for the PODs
1          The protocols used at the laboratories that have performed cancer bioassays of methanol,
2   particularly those of the ERF, differ from the more commonly used (e.g., NTP) protocols. The
3   unique features of the ERF study design and their implications to a methanol cancer risk
                                             5-37      DRAFT—DO NOT CITE OR QUOTE

-------
 1    assessment are discussed in Section 4.9.2.  Separate from these experimental design issues are
 2    considerations relative to the quality of the cancer bioassays and any associate uncertainties.
 3           The number of animals per dose group in ERF studies is often higher than the 50 animals
 4    per sex per dose group typically used in EPA and NTP studies, increasing the statistical power of
 5    the ERF cancer bioassay.  However, ERF sometimes shares controls across concurrent studies
 6    (Cruzan, 2009; Belpoggi et al., 1995). In contrast, EPA requires (U.S. EPA, 1998) and NTP
 7    generally uses (Melnick et al., 2007) concurrent, matched controls for each carcinogen bioassay.
 8    The use shared controls does not necessarily compromise a study, but the use of a concurrent,
 9    matched control is generally preferred as a means of further avoiding confounding factors and
10    increasing the reliability of a study regarding the interpretation of findings in treated animals.
11           The published report of the methanol bioassay (Soffritti et al., 2002a) indicates that the
12    experiment was performed according to good laboratory practice (GLP) and standard operating
13    procedures (SOP) of the ERF.  Further, an independent review of ERF (Huff, 2002) suggests that
14    quality control procedures associated with GLP were in place. However, questions have been
15    raised about the quality of studies at the ERF by European Food Safety Authority (EFSA, 2006)
16    in regards to the aspartame bioassays conducted by the ERF (Soffritti et al., 2006); by extension
17    this EFSA report has raised issues for consideration in regards to methanol. EFSA (2006) has
18    suggested that an inspection by the Italian GLP  compliance monitoring authority (Ministry of
19    Health) necessary to confirm GLP had not been conducted.87 The EFSA (2006) report also
20    identifies specific deviations from OECD guidelines (OECD, 2007), including a lack of a
21    complete analysis of the test substance, no clear information on the  stability of the substance, a
22    lack of clinical observations or macroscopic changes, a lack of hematological assays, a lack of
23    serology (e.g., to confirm  the presence of infection) and limited histopathology reports. While
24    these details may be recorded internally by the ERF as part of their standard protocol, because
25    there is no documentation of these details available for consideration, there remains some
26    uncertainty regarding the level at which they were performed.  There is limited evidence,
27    however, that these factors had a significant impact on the adequacy of the study for assessing
28    carcinogenic potential  (see Sections 4.2.1.3, 4.9.2 and 5.4.3.2).
29           EFSA (2006) also  expresses concern over the possibility of compromised pathological
30    diagnosis in the ERF aspartame study (Soffritti et al., 2006) due to extensive autolysis. ERF
31    performs pathological examinations on "dying animals undergoing necropsy" (Soffritti et al.,
32    2002a).  This creates difficulties in pathological examinations associated with cell autolysis that
33    can occur when pathology slides are prepared after natural  death.  The NTP (Hailey, 2004)
34    commented on the increased prevalence of autolysis in slides from the ERF (Soffritti et al., 2006)
35    aspartame study.  EPA conducted a detailed analysis of the  individual animal tissue data obtained
      87 Since the publication of the EFSA (2006) report, the EPA has confirmed through communication with the ERF
      laboratory (Knowles, 2008) that ERF is currently in the process of obtaining this certification.
                                                5-38      DRAFT—DO NOT CITE OR QUOTE

-------
 1    from ERF for their chronic methanol, MTBE, formaldehyde, and aspartame studies, and
 2    determined that autolysis and other causes of tissue loss did not substantially impact tissue
 3    denominators.  For most of the tissues evaluated there were more than 96 (individual animal)
 4    samples available for microscopic evaluation, and for all sites and dose groups, denominators
 5    were larger than for routine NTP bioassays (i.e., >50). Thus, missing tissues does not appear to
 6    have been a serious problem in the methanol study. While this analysis does not completely rule
 7    out the possibility that pathology slides and diagnoses were impacted by autolysis, it does
 8    indicate that this possibility would be offset by the large group size (response denominator)
 9    employed.  Further, even if autolysis was a confounding factor, its presence would not negate
10    positive cancer findings as autolysis would tend to decrease, not increase, the power to observe
11    an effect.
12           There were no differences in survival among the methanol dose groups of the Soffritti et
13    al. (2002a) study. However,  Cruzan (2009) has suggested that "While the survival at 104 weeks
14    was within the normal, but widespread, range for Sprague-Dawley rats, there was significant
15    early mortality  among all groups, including the controls" and that "the control group from an
16    inhalation study (Cruzan et al., 1998) had much better survival through 104 weeks than seen in
17    the RF methanol study."  Yet, according to Table 12 of the Cruzan (2009) article, 104-week
18    survival in male (-40%) and female (-50%) control rats of the Cruzan et al. (1998) study was
19    not discernibly different from the 104 week survival of male (-40%) and female (-50%) control
20    rats of the Soffritti et al. (2002a) methanol study (see Appendix E, Figures E-l and E-2).
21    Survival of male and female Sprague-Dawley rats in the Soffritti et al. (2002a) study at
22    104 weeks was greater than 40% in all but the female 5,000 ppm group. Further, at the NTP
23    (NTP,  2006), the 104-week survival of 353 control female Sprague-Dawley rats was 41.5%
24    (range of 28.3%-51% in 7 studies) using NTP's new  diet and  corn oil gavage, indicating that
25    survival in the Soffritti et al.  (2002a) methanol study was not low.
26          Studies  such as the Soffritti et al. (2002a) methanol study that allow test animals to live a
27    full life span can be difficult to interpret due to the need to distinguish between age-related and
28    chemical-related effects.  Full life-span studies may have advantages. Huff et al. (2009) note that
29    "studies truncated after 2 years of exposure do not allow sufficient latency periods for late-
30    developing tumors, such as the 80% of all human cancers that occur after 60 years of age."
31    Several recent publications have noted deficiencies with the 2-year study design used at the NTP
32    and have recommended extending the duration of rodent studies to increase the sensitivity of
33    their bioassays  (Huff et al., 2009; Huff et al., 2007; Maronpot et al., 2004; Bucher et al., 2002).
34          While arguments have been documented related to the possible confounding influence  of
35    infection and autolysis on the results obtained from the ERF, available evidence does not indicate
36    that these factors significantly influenced the observed lymphoma/leukemia response in the
37    methanol or other bioassays conducted at ERF.  In addition, for the purposes of this assessment

                                               5-39      DRAFT—DO NOT CITE OR QUOTE

-------
 1   and at the request of the EPA, the ERF and NEDO have provided additional study details beyond
 2   that which is normally available from published journal articles, including quality assurance
 3   reports and individual animal data.  Based on a review of this information, consideration of the
 4   issues, and absent additional data to the contrary, EPA has determined that both studies were
 5   sufficient for use in the assessment of risk from methanol exposure.
     5.4.3.2. Interpretation of Results of the Studies that are the Basis for thePODs
 6          There are a number of uncertainties regarding the interpretation of both the lymphoma
 7   response in male Sprague-Dawley rats (Soffritti et al., 2002a) that forms the basis of the oral
 8   CSF and the pheochromocytoma response in F344 rats (NEDO, 1985/2008b) that forms the basis
 9   for the inhalation CSF.
10          There is also a wide range in the background incidence of hemolymphoreticular tumors
11   reported in control groups of ERF studies. Between 1984 and 1997, incidence rates of
12   hemolymphoreticular neoplasms in control rats at ERF increased by 38% among male rats and
13   decreased by  12% among female rats (Caldwell et al., 2008).  Soffritti et al. (2007, 2006) reports
14   that among 2,265 untreated males and 2,274 untreated females the average incidence of
15   lymphomas and leukemias is 20.6% (range, 8.0-30.9%) in males and 13.3% (range, 4.0-25%) in
16   females. Caldwell et al. (2008) noted that for the  incidences of these lesions for the ERF colony
17   are relatively  low and stable across studies. EFSA (2006) and Cruzan (2009) consider it to be
18   high relative to other tumor types and relative to the background rate for this tumor type in
19   Sprague-Dawley rats from  other laboratories (see Section 4.9.2 for further discussion).
20          In the ERF bioassays, including methanol, hemolymphoreticular neoplasms were divided
21   into specific histological types (lymphoblastic lymphoma, lymphoblastic leukemia, lymphocytic
22   lymphoma, lympho-immunoblastic lymphoma, myeloid leukemia, histocytic sarcoma, and
23   monocytic leukemia) for identification purposes.  Upon examining slides from the aspartame
24   study conducted by the ERF, a PWG of the NTP at the NIEHS (Hailey, 2004) found that "The
25   diagnoses of lymphatic and histocytic neoplasms in the cases reviewed were generally
26   confirmed.  The NTP does not routinely subdivide lymphomas into specific histological types as
27   was done by the ERF, however the PWG accepted their more specific diagnosis if the lesion was
28   considered to be consistent with a neoplasm of lymphocytic, histocytic, monocytic, and/or
29   myeloid origin." The NTP PWG also noted that while lymphoblastic lymphomas, lymphocytic
30   lymphomas, lympho-immunoblastic lymphomas  and lymphoblastic leukemias as malignant
31   lymphomas can be combined, myeloid leukemias, histocytic sarcomas and monocytic leukemia
32   should be treated as separate malignancies and not combined with the other lymphomas for
33   statistical evaluation since they are of different cellular origin (Hailey, 2004).  Other researchers
34   have also noted this distinction between myeloid leukemias and histiocytic sarcomas and other
35   lymphomas (McConnell et al., 1986). To decrease the uncertainty in the combination of tumors
36   relied upon for dose-response modeling, the current dose-response modeling conducted for
                                              5-40      DRAFT—DO NOT CITE OR QUOTE

-------
 1   methanol did not include myeloid leukemia, histocytic sarcoma, monocytic leukemia, alone or in
 2   combination with lymphoblastic lymphoma, lymphoblastic leukemia, lymphocytic lymphoma,
 3   and lympho-immunoblastic lymphoma.
 4          As expressed by EFSA (2006) and others (Cruzan, 2009; Schoeb et al., 2009), there is
 5   concern that the lympho-immunoblastic lymphoma response in the ERF aspartame study
 6   (Soffritti et al., 2006, 2005) was caused by or confused with sequalae of aM pulmonis infection.
 7   Infection of the ERF colony with M. pulmonis has not been confirmed (Caldwell et al., 2008)
 8   and, as discussed in Section 4.9.2, a link between M. pulmonis infection and induction of
 9   lymphoma in rats has not been established in the literature. As noted in Section 4.9.2, there is
10   evidence suggesting that respiratory infections may have confounded the interpretation of lung
11   lesions in ERF studies. Lymphoma illustrations in 2 ERF studies (Figure 10 of Soffritti et al.
12   [2005] and Figures 1-5 of Belpoggi et al. [1999]), suggest that ERF MTBE and aspartame
13   bioassays may have been confounded by a respiratory infection such as M pulmonis and that
14   lesions associated with this infection may have been interpreted as lymphoma (Schoeb et al.,
15   2009). However, other ERF lymphoma diagnoses in multiple rat organ systems, including the
16   lung, have been confirmed by an independent working group panel of six NIEHS pathologists
17   (Hailey, 2004).  In addition, the incidence of "lung-only" lympho-immunoblastic lymphomas
18   was evenly distributed across the control and 0, 500 5000, 20,000 ppm dose groups for male (9,
19   10, 14 and 13) and female (3, 5, 6 and 7) rats of the Soffritti et al. (2002a) methanol study
20   (Schoeb et al., 2009). Consequently, removing "lung-only" lympho-immunoblastic lymphomas
21   from consideration and using only lymphomas from organ systems not likely to be confounded
22   by a respiratory infection (i.e.,  subtracting the lung-only incidence from lympho-immunoblastic
23   lymphomas reported in Table 5-6) still results in significant dose-response curves for this lesion
24   with p values of 0.20 and 0.42  for males and females, respectively (see Figure 5-6). The
25   BMDLio estimates  of 99 (males) and 110 (females) mg metabolized methanol/day are about 50%
26   higher than metabolized methanol BMDLio estimates for this endpoint when "lung-only"
27   responses are included (Appendix E, Table E-7).
28
29
                                              5-41      DRAFT—DO NOT CITE OR QUOTE

-------
                          Multistage Cancer Model with 0.95 Confidence Level
                                                                           ;r Model with 0.95 Confidence Level
0.35
0.3
0.25
0.2
0.15
0.1
0.05

Multistage Cancer — ' ' -
Males (p= 0.20)



-^

^^^^___^— —
^-~^^

^^^^^





^^—~~



BHDL BHD
                                                       1
                                                                         Multistage Cancer
                                                                         Linear extrapolation
                                                                     Females (p=
.42)
              08:35 09/04 2009
                                   100
                                   dose
                                                        08:40 09/04 2009
          Figure 5-6. Lympho-immunoblastic Lymphoma minus "lung-only" response in rats of
          Soffritti et al. (2002a) methanol study versus methanol metabolized (mg/day).
 2           For increases in 2 other tumor types (ear duct and head/oral cavity tumors) reported in the
 3    ERF methanol study (Table 4-2), an independent review of the 75 pathology slides from the ERF
 4    aspartame study has suggested differences in interpretation. After reviewing these slides, the
 5    NTP PWG noted that "about half of hyperplastic and neoplastic lesions in the ear duct or oral
 6    cavity were more severely classified by ERF study pathologists, compared with diagnosis from
 7    the PWG (EFSA, 2006).  Though a similar review has not been conducted of the Soffritti et al.
 8    (2002a) ERF methanol bioassay results, there is uncertainty regarding the ERF interpretation of
 9    these lesions.  For this reason, these lesions were not considered in the derivation of the oral
10    CSF.
11           Another uncertainty that confounds the interpretation of some ERF studies is the
12    possibility of litter effects in ERF test groups.  Bucher (2002) has reported that ERF does not
13    randomize the assignment of animals to treatment groups, but generally "assigns all animals
14    from a given litter to the same treatment group, recording which litter each animal came from."
15    This approach would make it more difficult to distinguish the relative contributions of the
16    chemical and genetics to the effect of concern.  In response to an EPA query regarding this matter
17    (Knowles, 2008), ERF has stated that "the assignment of test animals to dose groups will vary in
18    ERF studies according to the experimental protocol and aims of the research" and "in the case of
19    experiments in which exposure begins at 6-8 weeks of age (for example BT960, methanol),
20    randomization is performed so as to have no more than one female and one male from each litter
21    in each experimental group." For other experiments in which exposure begins during prenatal
22    life,88 "randomization is performed on the breeders," but the offspring are not randomized across
23    dose groups in order to "..simulate as much as possible the human situation in which all
24    descendents are part of a population."
      88 For some chemicals such as vinyl chloride (Maltoni and Cotti, 1988) and aspartame (Soffriti et al., 2006), ERF
      has started exposure as early as in utero, This early exposure study design can markedly increase the sensitivity of a
      cancer bioassay (Maltoni and Cotti, 1988; Soffriti et al., 2006; Melnick et al., 2007).
                                                5-42      DRAFT—DO NOT CITE OR QUOTE

-------
 1          There is uncertainty regarding the pheochromocytoma response observed in the NEDO
 2    (1985/2008b) study with respect to both its relation to exposure and to its pathology. As
 3    discussed in Section 4.2.2.3, the incidence of pheochromocytomas in female rats exhibited a
 4    dose-response trend (Cochrane-Armitage/? < 0.05).  While the incidence of 13.7% (7/51) in the
 5    high-dose group was significantly elevated (p < 0.05) over NTP historical controls, it was not
 6    significantly elevated over the concurrent control rate of 4% (2/50). Much of the uncertainty is
 7    inherent in difficulties associated with the characterization of this lesion. According to NTP
 8    (2000), the primary criterion used to distinguish pheochromocytomas from nonneoplastic adrenal
 9    medullary hyperplasia, the presence of mild-to-moderate compression of the adjacent tissue, can
10    be a difficult determination. Further, while NEDO (1985/2008b) reported adrenal effects as
11    "hyperplasia of medullary cells" and "pheochromocytoma," NTP pathologists categorize
12    pheochromocytomas into three types: benign, complex and malignant (NTP, 2007, 1999).  This
13    is an important distinction as complex and malignant pheochromocytomas are a rare tumor type,
14    occurring spontaneously in female F344 rats at rates ranging from 0.1% to 0.7% (NTP, 2007,
15    1999; Haseman et al., 1998) and with cell proliferation activity that is much higher than benign
16    pheochromocytomas (Pace et al., 2002).  Thus, severity of the pheochromocytoma response
17    reported by NEDO (1985/2008b) is uncertain, potentially ranging from mischaracterized
18    hyperplasia to highly proliferative and potentially metastatic malignancies. Finally, the NEDO
19    study was a two-year study, and these lesions, which include diffuse hyperplasia, nodular
20    hyperplasia, and pheochromocytoma, progress with  age. Thus, it is possible that a life-span
21    study would have detected a more severe carcinogenic response (e.g., progression of the mid-
22    dose group hyperplastic responses as reported in Table 4-5).
      5.4.3.3. Consistency across ChronicBioassays for Methanol
23          The observation of a lymphoma response in rats (Soffritti et al., 2002a) and mice (Apaja,
24    1980), along with reported associations between lymphomas and human exposure to methanol's
25    metabolite, formaldehyde (see Section 4.9), contributes significantly to the cancer weight-of-
26    evidence determination. This was the only carcinogenic response that was observed in more than
27    one animal bioassay. However, tissue concordance across species, strains and routes of exposure
28    is not assumed, and a lack of such concordance does not negate the validity of individual
29    findings nor the potential relevance of such findings to humans.
30          Besides the Soffritti et al. (2002a) drinking water study of rats (2 years) reported by ERF,
31    the only other chronic studies available are the Apaja (1980) lifespan drinking water study study
32    in Swiss mice which did not include a concurrent untreated control group, and those reported in
33    the Japanese NEDO (1987, 1985/2008a, 2008b) study of monkeys (7-29 months), mice
34    (18 months), and F344 rats (2 years).  None of the NEDO studies involved lifetime evaluations,
35    and only the rat study can be said to cover a significant portion of the test species life span. The
36    only organ system that exhibited an increase in effects in both inhalation and drinking water
                                               5-43       DRAFT—DO NOT CITE OR QUOTE

-------
 1    studies was the testes. High-dose rats of the Soffritti et al. (2002a) methanol study exhibited an
 2    increase in testicular interstitial cell adenomas, and the incidence of testicular hyperplasia was
 3    increased in high-dose rats of the NEDO (1987, 1985/2008b) study.  However, neither effect was
 4    statistically increased over controls, and both effects were well within their historical control
 5    ranges. An increase in lymphomas was the only carcinogenic effect reported in the Soffritti et al.
 6    (2002a) and Apaja (1980) drinking water studies that is considered dose related and quantifiable,
 7    and male pulmonary adenoma/adenocarcinoma and female pheochromocytoma were the only
 8    carcinogenic effects from the NEDO (1987, 1985/2008b) inhalation study that are considered
 9    exposure related and quantifiable.
10          As discussed in Section 4.9.2, there are several differences between the NEDO (1987,
11    1985/2008b) and Soffritti et al. (2002a) studies conducted in rats which limit their direct
12    comparison and may explain some of the differences in response. These include route of
13    exposure, lifetime evaluation period, and strain of species used. Pulmonary adenomas observed
14    in the NEDO inhalation study could be portal-of-entry effects associated with the inhalation
15    route of exposure.  Differences in systemic effects observed in the two studies such as the
16    pheochromocytoma response in the NEDO (1987, 1985/2008b) study and the lymphoma
17    responses observed in the Soffritti  et al. (2002a) study are systemic responses and differences
18    would not be expected based on route of exposure. Differences in the evaluation period between
19    the two studies may have contributed to the lack of endpoint concordance.  In the Soffritti et al.
20    (2002a) study, the animals were administered methanol for 104 weeks and then followed until
21    the completion of their natural lifetime. The average life span for these animals was 94 and 96
22    weeks for males and females, respectively, with animals living as long as 153 weeks (female).
23    However, this does not explain the difference in lymphoma response between the studies as
24    many of the lympho-immunoblastic lymphomas (most common type observed) were observed
25    prior to 104 weeks (control - 5/9; 500 ppm - 7/17; 5,000 ppm - 13/19; 20,000 ppm - 11/21). The
26    NEDO (1987, 1985/2008b) study was conducted in F344 rats versus the  Sprague-Dawley rats
27    used in the Soffritti et al. (2002a) ERF  study. More importantly, the background rates of selected
28    types of lymphomas and leukemias are very different between the two strains of rats. In the
29    F344 rat, there is a high background rate of mononuclear cell leukemia, while there is a much
30    lower background rate of this leukemia type in the Sprague-Dawley rat (Caldwell et al., 2008).
31    The types of lymphoma reported in the Sprague-Dawley rat by Soffritti et al. (2002a) following
32    methanol administration are rarely diagnosed in the F344 rat. Thus, the strain difference
33    between the two studies is a likely  explanation for the fact that lymphomas were only increased
34    in the Soffritti et al. (2002a) rat study. NTP (2007, 1999) reports do not suggest a significant
35    difference between F344 and Sprague-Dawley rats with respect to pulmonary adenomas and
36    pheochromocytomas. Another possible explanation for this difference includes a different
37    profile of circulating metabolites associated with oral first-pass liver metabolism.

                                               5-44      DRAFT—DO  NOT CITE OR QUOTE

-------
      5.4.3.4. Ch oice of En dp oin t for POD Deriva tion
 1          Keeping in mind the aforementioned uncertainties associated with the interpretation of
 2    the Soffritti et al.  (2002a) and NEDO (1985/2008b) study results, the choice of tumors to use for
 3    the derivation of the oral and inhalation CSFs was made on the basis of the appearance of a dose-
 4    related increase in response rates for the selected tumor categories.  Analysis of lymphomas used
 5    a pair-wise statistical comparison (Fisher's Exact test) and a trend test (Cochran Armitage trend
 6    test) to determine whether the slope of that trend was statistically significantly greater than 0.
 7    The Fisher's Exact result for the male rat high-dose group and the results of the Cochran
 8    Armitage Trend Test were both/? < 0.01. The decision not to include the liver tumors in the
 9    dose-response analysis was made based on a lack of dose response according to pair-wise and a
10    trend test results versus concurrent controls This is not to say that the slight increase in the
11    incidence of this tumor type observed in all dose groups of male rats was not related to methanol
12    exposure (this is a relatively rare Sprague-Dawley rat tumor),  only that the increase was not
13    statistically significant and did not contribute significantly to the overall tumor response.
14          As discussed in Sections 4.2.2.3 and 4.9.2, the high-dose incidences for pulmonary
15    adenomas/adenocarcinomas were increased over concurrent controls (p < 0.05).  While the high-
16    dose incidences of pheochromocytomas in the NEDO (1985/2008b) study were not statistically
17    increased over concurrent  controls, the dose-response for both tumor types represents increasing
18    trends (Cochran Armitage  trend test; p < 0.05), and  in both cases, the high-dose response
19    incidences were considerably elevated over historical  control incidences (p < 0.05) within their
20    respective sex and strain.  Further, both tumor responses are accompanied by proliferative
21    changes (e.g., hyperplastic responses) in their respective cell types that suggest tumor
22    progression.
      5.4.3.5. Ch oice ofSp ecies/Gen der
23          The oral CSF was based on male rat lymphomas rather than female rat lymphomas.  The
24    inhalation IUR was based  on female rat pheochromocytomas rather than male rat
25    adenoma/carcinomas.  In both cases, the gender that exhibited the steeper dose response and the
26    higher risk estimate was chosen.
27          Both the CSF and IUR were based on rat studies. Use of the Apaja (1989) mouse data
28    would have resulted in a 5-fold higher, but less reliable oral CSF due to a high level of
29    uncertainty associated with the Apaja (1989) study,  which contained limited experimental detail
30    and did not include a concurrent control group (see  Section 4.2.1.3).
      5.4.3.6. Ch oice of Mod el for POD Deriva tion
31          The multistage-cancer model contained in EPA's BMDS version 2.1 was used to derive
32    both the CSF and IUR  estimates for methanol. When no biologically-based cancer model exists
33    and evidence for a nonlinear cancer MOA is lacking, as is the case for methanol, the preference

                                               5-45      DRAFT—DO NOT CITE OR QUOTE

-------
 1    within the EPA's IRIS program is for the use of a multistage model. There is uncertainty
 2    associated with whether the multistage model is the most appropriate choice, but in the absence
 3    of a biologically based model, dose-response modeling is largely a curve-fitting exercise, and the
 4    multistage model is sufficiently flexible for most cancer bioassay data. In the case of the oral
 5    CSF, individual animal response data was obtained from the authors of the principal study
 6    (Sofffritti et al., 2002a) and a multistage-Weibull time-to-tumor model was applied to determine
 7    whether the lifespan study design of the study had an appreciable impact on  the dose-response
 8    analysis.  As described in Appendix E, time-to-tumor modeling and multistage quantal modeling
 9    gave similar results, and the tumor responses modeled did not exhibit significant time
10    dependence on dose.
11
      5.4.3.7. Ch oice of Dose Metric
12          The total methanol metabolized was selected over AUC or the Cmax as the most
13    appropriate dose metric for derivation of the oral cancer slope factor and inhalation unit risk
14    primarily because it provided the best fit to response data, particularly lymphoma incidence from
15    Soffritti et al. (2002a) (see Figures 5-3 through 5-5 and Table E-7 of Appendix E) and Apaja
16    (1989) (see Table E-17 of Appendix E). Also, lymphomas and respiratory effects have been
17    observed in studies conducted with formaldehyde, and lymphomas have been observed in
18    chronic bioassays conducted with other compounds that convert to formaldehyde (i.e., MTBE
19    and aspartame).  As discussed in Section 4.9.3, metabolites of methanol, particularly
20    formaldehyde, may play a role in the MOA.
21          In considering the dose-response  relationship for methanol-induced carcinogenesis, a key
22    factor is the saturation of metabolism since metabolic transformation to formaldehyde and
23    generation of oxidative stress are considered likely candidates in the mode of action.  Cruzan
24    (2009) indicates that saturation occurs at dose of "600 mg/kg," but saturation depends on the
25    dose rate, not the total administered dose. For example, if 600 mg/kg is given in a single bolus,
26    the internal concentration immediately following that bolus could well be high enough to
27    saturate metabolism, while the same total dose ingested over the course of a  day might not.
28          To aid in interpretation of the Soffretti et al.  (2002a) bioassay, water ingestion in rats was
29    assumed to shift between nocturnal (high activity) and diurnal (low activity) periods, each lasting
30    12 hours. Rats were assumed to consume 20% of their daily water ingestion during the diurnal
31    period and 80% during the nocturnal period. Ingestion in each period was assumed to occur in
32    "bouts" which were treated as periods of continuous (zero-order) infusion to the stomach.
33    During the nocturnal period each bout was assumed to last 45  minutes, followed by 45 minutes
34    without ingestion (overall  period is 1.5 hours) and during the diurnal period  the bout was
35    assumed to last on 3 hours followed by 2.5 hours without ingestion (overall period is 3 hours).

                                               5-46       DRAFT—DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
10
11
12
13
14
15

16
17
18
19
20
21
       Given this exposure pattern, the total amount metabolized per day (after periodicity is
reached) in a 420 g rat (average weight in Soffretti et al., 2002a) was estimated using the PBPK
model, with the results shown in Figure 5-7. The amount increases almost linearly with
exposure until ~ 400 mg/kg/d, but continues to increase above that point, becoming almost
completely saturated by 2,000 mg/kg/d. This pattern occurs in part because of the circadian
ingestion pattern. The more rapid ingestion rate during the dark cycle leads to the highest
internal concentrations and hence the initial metabolic saturation during that part of the day.  But
the lower ingestion (light) period, internal concentrations drop, allowing for an exposure range
(400-1600 mg/kg/d) where nocturnal metabolism is  saturated but diurnal metabolism is not.
           ISO-
              cn
             T)
              o
             "
           160-

           140-

           120-

           100

            80-
              0)
              £  60-
             !s
              O  40-1
                 20-
                                          Total methanol  metabo ism
                                         from drinking water exposure
                         200
                               400
                                     600
                                           800
                                                 1,000  1,200   1,400   1,600  1,800  2,000  2,200
                                             Ingestion (mg/kg/d)
    Figure 5-7. Total amount metabolized per day (after periodicity is reached) in a 420 g
    rat
       While, based on this exposure-dose pattern, one might expect a similar exposure-
response relationship, this pattern does not include detoxification mechanisms.  If such
mechanisms also saturate, then it is possible for the slower increase in total metabolism above
400 mg/kg/d to result in a significant increase in effect, though full metabolic saturation at
-2,000 mg/kg/d would still be expected to result in a maximal effect at that exposure level.
5.4.3.8. Ch oice ofAnim al-to-Hum an Ext rap ola tion Meth od
       A PBPK model was used to extrapolate animal-to-human concentrations. The estimated
methanol metabolized for each dose administered to the animals in NEDO (1985/2008b) and
Soffritti et al. (2002a) were determined using the animal PBPK model, and then the BMDLio
determined by the methods described previously (Section 5.4.2.1).  Assuming that metabolized
methanol distributes in the body according to body weight to the 3/4 power, the rat mg-day
BMDLio was converted to a human mg-day BMDLio.  The human PBPK model (Appendix B)
                                               5-47
                                                   DRAFT—DO NOT CITE OR QUOTE

-------
 1   was then used to convert this human mg-day values for total methanol metabolized back to a
 2   human equivalent methanol oral dose HED(BMDLio) of 51.5 mg/kg-day for lymphomas in the
 3   male rat, and a human equivalent methanol inhalation concentration HEC(BMCLio) of
 4   81.9 mg/m3 for pheochromocytomas in the female rat. If traditional dosimetry assumptions are
 5   used, the HED(BMDLio) and HEC(BMCLio) estimates would have been approximately 4-fold
 6   higher than the value derived using the PBPK model.
 7          As discussed in Sections 3.4 and 5.3.4, the PBPK models do not describe or account for
 8   background levels of methanol, formaldehyde or formate, and background levels were subtracted
 9   from the reported data before use in model fitting or validation (if not already  subtracted by
10   study authors), as described below. This approach was taken because the relationship between
11   background doses and background responses is not known, because the primary purpose of this
12   assessment is for the determination of noncancer and cancer risk associated with increases in the
13   levels of methanol or its metabolites (e.g., formate, formaldehyde) over background, and because
14   the subtraction of background levels is not expected to have a significant impact on PBPK model
15   parameter estimates as background levels of methanol and its metabolites are low relative to
16   exposure levels used in methanol bioassays. Further, while it is possible that background levels
17   of methanol or its metabolites contribute to background responses for some adverse effects, the
18   results of dose-response modeling of cancer endpoints using "background dose" models suggest
19   that this contribution is relatively small (see discussion in Appendix E, Section E.4).
20
     5.4.3.9. Human Relevance of Cancer Responses Observed in Rats and Mice
21          As discussed in Sections 4.9.2, there is human evidence for the association of lymphomas
22   with a metabolite of methanol, formaldehyde.  However, there is no information available in the
23   literature regarding the observation of cancer in humans following chronic administration of
24   methanol.  The only observations in animals were noted in the chronic studies of methanol
25   conducted by Apaja (1980), Soffritti et al. (2002a) and NEDO (1985/2008b) and there is
26   uncertainty associated with the interpretation of the tumor responses reported in these studies.
27   As a consequence, the overall WOE, while convincing, is not strong.
                                              5-48       DRAFT—DO NOT CITE OR QUOTE

-------
      6. MAJOR CONCLUSIONS IN THE CHARACTERIZATION OF HAZARD AND DOSE
                                           RESPONSE
      6.1. HUMAN HAZARD POTENTIAL
 1          Methanol is the smallest member of the family of aliphatic alcohols. Also known as
 2    methyl alcohol or wood alcohol, among other synonyms, it is a colorless, very volatile, and
 3    flammable liquid that is widely used as a solvent in many commercial and consumer products.  It
 4    is freely miscible with water and other short-chain aliphatic alcohols but has little tendency to
 5    distribute into lipophilic media. Methanol can be formed in the mammalian organism as a
 6    metabolic byproduct and can be ingested with foodstuffs, such as fruits or vegetables.  A
 7    potential for human exposure exists today in the form of the artificial sweetener, aspartame,
 8    which is a methyl ester of the dipeptide aspartyl-phenylalanine. Methanol is the major anti-
 9    freeze constituent of windshield washer fluid. Its use as a fuel additive for internal combustion
10    engines is, as yet, limited by its corrosive properties.
11          Because of its very low oil:water partition coefficient, methanol is taken up efficiently by
12    the lung or the intestinal tract and distributes freely in body water without any tendency to
13    accumulate in fatty tissues.  It can be metabolized completely to CO2, but may also, as a regular
14    byproduct of metabolism, enter the Ci-pool and become incorporated into biomolecules. Animal
15    studies indicate that blood methanol  levels increase with the breathing rate and that metabolism
16    becomes saturated at high exposure levels.  Because of its volatility it can also be excreted
17    unchanged via urine or exhaled air.
18          The acute toxicity in laboratory animals in response to high levels of exposure results
19    from CNS depression.  NEDO (1987) reported that methanol blood levels around 5,000 mg/L
20    were necessary to cause clinical signs and CNS changes in monkeys. In humans, however, acute
21    toxicity is caused by metabolic acidosis that appears to affect predominantly the nervous system,
22    with potentially lasting effects such as blindness, Parkinson-like symptoms, and cognitive
23    impairment. These effects can be observed in humans when blood methanol levels exceed
24    200 mg/L. The species differences in toxicity from acute exposures appear to be the result of a
25    limited ability of humans to metabolize formic acid.
26          Despite the existence of many case reports on acute human exposures, the knowledge
27    base for long-term, low-level exposure of humans to methanol is limited. The current TLV for
28    methanol is 200 ppm (262 mg/m3) (ACGIH, 2000). Controlled experiments with human
29    volunteers indicate that only minor neurobehavioral changes occur following 4-hour exposure to
30    this concentration. A limited study on self-reported health effects in 66 persons exposed to
31    methanol at levels that  came close to or exceeded the NIOSH short-term ceiling of 800 ppm
32    (1048 mg/m3), in comparison with an age-matched group of 66 less or not exposed persons,
33    suggested a statistically significant increase in the incidence of CNS-related symptoms, such as
34    dizziness, nausea, headache, and blurred vision (Frederick et al., 1984). Impaired vision and
                                               6-1      DRAFT -DO NOT CITE OR QUOTE

-------
 1    nasal irritation were observed in a study of 33 methanol-exposed workers (Kawai et al., 1991).
 2    None of the case reports or human studies have investigated cancer as a potential outcome of
 3    methanol exposure.
 4          A number of reproductive, developmental, subchronic and chronic exposure duration
 5    studies have been conducted in mice, rats, and monkeys. This summary will focus primarily on
 6    reproductive and developmental toxicity, and cancer as the main endpoints of concern. Sections
 7    4.7, 5.1.1 and 5.2.1 contain more extensive summaries that consider the dose-related effects that
 8    have been observed in other organ systems following subchronic or chronic exposure.
 9          Although there is no evidence in humans, methanol has shown to be a reproductive and
10    developmental toxicant in several animal studies.  No studies have been reported in which
11    humans have been exposed subchronically or chronically to methanol by the oral route of
12    exposure, and thus would be suitable for derivation of an oral RfD. Data exist regarding effects
13    from oral exposure in experimental animals, but they are more limited than data from the
14    inhalation route of exposure (see Sections 4.2, 4.3, and 4.4). Two oral studies in rats (Soffritti
15    et al., 2002a; U.S. EPA,  1986), one oral study in mice (Apaja, 1980) and several inhalation
16    studies in monkeys, rats and mice (NEDO, 1987, 1985/2008a, 2008b) of 90-days duration or
17    longer have been reported. While some noncancer effects of methanol  exposure were noted in
18    these studies, principally in the liver and brain, they were either not quantifiable due to study
19    limitations or occurred at high doses relative to reproductive/developmental effects. As
20    discussed below, the results of inhalation reproductive/developmental toxicity studies in rats
21    (NEDO, 1987), mice (Rogers et al., 1993a),  and monkeys (Burbacher et al., 2004a, 2004b,
22    1999a, 1999b) are the principal considerations for both the RfD and RfC values derived in this
23    assessment.
24          A larger number of studies have used the inhalation route to assess the potential of
25    reproductive or developmental toxicity of methanol in mice, rats, and monkeys,  with
26    concentrations ranging from 200 to 20,000 ppm (blood levels reaching as high as 8.65 mg/mL).
27    To sum up the findings, rat dams survived even the highest doses without gross signs of toxicity,
28    but their offspring were severely affected (Nelson et al.,  1985).  Two more inhalation studies,
29    Rogers et al. (1993a,1993b) and Rogers and Mole (1997), confirmed that methanol causes
30    exencephaly and cleft palate in mice, the most sensitive days being GD6 and GD7 (i.e., early
31    organogenesis). These severe malformations were observed at  exposure concentrations of
32    5,000 ppm or above. Nelson et al. (1985) and Rogers et al. (1993a) also observed an increased
33    occurrence of ossification disturbances and skeletal anomalies at methanol concentrations
34    > 2,000 ppm, of which cervical ribs in mouse fetuses is considered the critical effect for toxicity
35    value derivation in this review. A study conducted in pregnant cynomolgus monkeys that were
36    exposed to 200-600 ppm methanol for 2.5 hours/day throughout premating, mating, and
37    gestation showed no signs of maternal or fetal toxicity.  The potential compound-related effects

                                                6-2      DRAFT—DO NOT CITE OR QUOTE

-------
 1    noted were a shortening of the gestation period by less than 5% and developmental neurotoxicity,
 2    particularly delayed sensorimotor development (Burbacher et al., 2004a, 2004b, 1999a, 1999b).
 3          While all of the above studies were conducted with exposure durations of 7 hours/day or
 4    less, NEDO (1987) conducted a series of developmental/reproductive studies in rats that used
 5    exposure times of 20 hours/day or more at concentrations between 500 and 5,000 ppm. Atwo-
 6    generation study by these researchers that exposed the dams throughout pregnancy and the pups
 7    through 8 weeks of age, demonstrated dose-dependent reductions in brain weights that forms the
 8    basis for the RfC derived in this review.
 9          Carcinogenic effects following methanol exposure were observed in a chronic drinking
10    water study in Eppley Swiss Webster mice (Apaja, 1980) and two chronic rat studies: a drinking
11    water study of Sprague-Dawley rats (Soffritti et al., 2002a) and an  inhalation study of F344 rats
12    (NEDO, 1985/2008b).  Following administration via drinking water, both Apaja (1980) and
13    Soffritti et al. (2002a) observed positive dose-response trends for increases in the incidence of
14    lymphomas in both test animal genders. Soffritti et al. (2002a) characterized the lymphomas in
15    their study as lymphoreticular, principally lympho-immunoblastic.  EPA re-analyzed the
16    lymphoma data from the Soffritti et al. (2002a) study for quantification purposes, combining
17    only tumors of the same cell type origin. There was a slight increase in hepatocellular
18    carcinomas in male rats of all exposure groups of this study that was not statistically elevated
19    over controls in any group, but potentially this tumor is related to methanol exposure given the
20    low historical background rate for this tumor in this rat strain. As discussed in Section 5.4.1.1,
21    the other tumor increases reported by Soffritti et al. (2002a) are not quantifiable or were
22    considered hyperplastic rather than carcinogenic following a review by NTP pathologists (EFSA,
23    2006; Hailey, 2004).  No tumor responses were significantly increased over controls in the
24    chronic inhalation bioassays performed by NEDO (1987) in monkeys, and mice, but the high-
25    dose incidences for pulmonary adenomas/adenocarcinomas in male rats was elevated over
26    concurrent controls and pheochromocytomas in female rats were significantly elevated over
27    historical control incidences for these tumor types within their respective sex and strain. The
28    dose response for both of these tumor types represents increasing trends (Cochran Armitage
29    trend test; p < 0.05).  Further, both tumor responses are accompanied by proliferative changes
30    (e.g., hyperplastic responses) in their respective cell types.

      6.2. DOSE RESPONSE
      6.2.1. Noncancer/Inhalation
31          Clearly defined toxic endpoints at moderate exposure levels have been observed only in
32    reproductive and developmental toxicity studies. Three endpoints  from developmental toxicity
33    studies were considered for derivation of the RfC: formation of cervical ribs in CD-I mice
34    exposed to methanol during organogenesis (Rogers et al., 1993a), deficits in sensorimotor
                                               6-3       DRAFT—DO NOT CITE OR QUOTE

-------
 1    development as measured by VDR tests administered to monkeys exposed to methanol
 2    (Burbacher et al., 2004a, 2004b, 1999a, 1999b), and reduced brain weights in rats exposed to
 3    methanol from early gestation through 8 weeks of postnatal life (NEDO, 1987). For the purpose
 4    of comparability and to better illustrate methodological uncertainty, reference values were
 5    derived for all of these endpoints using a BMD modeling approach which evaluated several
 6    models and various measures of risk. In the present review, mostly because of a paucity of
 7    adequate long-term or developmental oral studies and the existence of several inhalation studies
 8    that examined sensitive subpopulations (pregnant mothers, developing fetuses and neonates) in
 9    various species, it was decided to use the critical effect from an inhalation  study to derive an
10    RfD.  Thus, the criteria and rationales on which the RfC assessment is based also form the basis
11    for the RfD derivation.
12          The Rogers et al. (1993a) inhalation study is a multidose developmental study that was
13    considered for use in the derivation of a reference value.  The exposure concentrations in this
14    study were 0, 1,000, 2,000, and 5,000 ppm administered for 7 hours/day on GD7-GD17. The
15    BMD evaluation, based on the nested log-logistic model of BMDS version 2.1, produced
16    BMD/BMDL values in terms of internal peak blood methanol (Cmax).  PBPK modeling was used
17    to convert the internal animal dose metrics to HECs,  and a UF of 100  was  applied to yield RfCs
18    of 10.4 mg/m3 and 13.6 mg/m3 for 5 and 10% extra risk, respectively.
19          Reproductive and developmental neurobehavioral effects observed in monkeys following
20    methanol inhalation exposure (Burbacher et al., 2004a, 2004b,  1999a, 1999b ) were also
21    considered for use in the derivation of a reference value. M. fascicularis monkeys were exposed
22    to 0, 262, 786, and 2,359 mg/m3 methanol 2.5 hours/day, 7 days/week during premating/mating
23    and throughout gestation (approximately 168 days).  Delayed sensorimotor development as
24    measured by  a VDR test was the only effect in this study that exhibited a dose-response and is a
25    measure of a  functional deficit that is consistent with early developmental  CNS effects (e.g.,
26    brain weight changes) that have  been observed  in rats (NEDO,  1987). Though there is
27    uncertainty associated with this effect and its relation to  methanol exposure, a BMD analysis was
28    performed for comparative purposes. BMD/BMDL values for the VDR endpoint were estimated
29    using AUCs derived from a monkey PBPK model of blood methanol data reported in the
30    Burbacher et al. (1999a) study. A human methanol PBPK modeling was then used to convert the
31    internal AUC BMDL to an FIEC, and a UF of 100 was applied to yield a reference value estimate
32    of 1.7 mg/m3.
33          Reduced brain weight was evaluated based on the results of a two-generation study by
34    NEDO (1987) in which fetal  rats and their dams were exposed  from the first day  of gestation
35    until 8 weeks of age, and brain weights were determined at 3, 6, and 8 weeks of age. To obtain
36    reference value estimates from these studies, a rat PBPK model was used to predict PODs in
37    terms of internal doses, which were divided by  UFs and  converted to HEC reference values via a

                                               6-4       DRAFT—DO NOT CITE OR QUOTE

-------
 1    human PBPK model (see Table 5-4).  BMD modeling was executed using two different BMRs,
 2    one S.D. (as is usual with continuous data) and 5% relative (to control response) risk. The
 3    resulting reference value estimates were 2.4 and 1.8 mg/m3 (5% relative risk and 1 S.D.,
 4    respectively) for reduced brain weight at 6 weeks of age following gestational and postnatal
 5    exposure.
 6          Despite the variety of approaches, different critical effects, and different data sources, all
 7    reference value estimates fell within a narrow range. The reference value associated with the
 8    BMD estimate of the dose corresponding to a one S.D. decrease in brain weight in male rats at 6
 9    weeks post-birth observed in the NEDO (1987) developmental toxicity study is considered most
10    suitable for derivation of the methanol chronic RfC due to the relevance of the exposure
11    scenario/study design (see Sections 5.1.2.2 and 5.3), and endpoint (see Section 5.3) to the
12    potential for developmental effects in neonatal humans, the relative robustness of the dose-
13    response data and because it resulted in one of the lowest reference values of the BMD
14    derivations (see Table 5-4). Thus, the proposed chronic RfC for exposure to methanol is 2
15    mg/m3, an evaluation that includes a UFH of 10 for intraspecies variability, a UFA of 3 to address
16    the pharmacodynamic component of interspecies variability, and a UFD of 3 for database
17    uncertainty.
18          The confidence in this RfC is medium to high. Confidence in the NEDO (1987)
19    developmental studies is medium to high. While there are issues with the lack of reporting
20    detail, the critical effect (brain weight reduction) has been reproduced in an oral study of adult
21    rats (U.S. EPA, 1986c),  and the exposure regimen involving pre- and postnatal exposures
22    addresses a potentially sensitive human  subpopulation.  Confidence in the database is medium.
23    Despite the fact that skeletal and brain effects have been demonstrated and corroborated in
24    multiple animal studies in rats, mice, and monkeys, some study results were not quantifiable,
25    there is uncertainty regarding which is the most relevant test species, and there is limited data
26    regarding reproductive or developmental toxicity of methanol in humans. There is also
27    uncertainty regarding the potential active agent—the parent compound, methanol, formaldehyde,
28    or formic acid.  There are deficiencies in our knowledge of the metabolic pathways of methanol
29    in the human fetus during early organogenesis, when the critical effects can be induced  in
30    animals. Thus, the medium-to-high confidence in the critical study and the medium confidence
31    in the database together warrant an overall confidence descriptor of medium to high.

      6.2.2. Noncancer/Oral
32          There is a paucity of scientific data regarding the outcomes of chronic oral exposure to
33    methanol.  No data exist for long-term methanol exposure of humans. A subchronic (90-day)
34    oral study in  Sprague-Dawley rats reported brain and liver weight changes, with some evidence
35    for minor liver damage at 2,500 mg/kg-day that was not supported by histopathologic findings
                                                6-5       DRAFT—DO NOT CITE OR QUOTE

-------
 1    (U.S. EPA, 1986c). Liver necrosis was reported in Eppley Swiss Webster mice that consumed
 2    approximately 2000 mg/kg-day (Apaja, 1980).  In the only other study that administered
 3    methanol chronically to animals by the oral route, Soffritti et al. (2002a) reported that, overall,
 4    there was no pattern of compound-related clinical signs of toxicity in Sprague-Dawley rats
 5    exposed to up to approximately 2,000 mg/kg-day. The authors further reported that there were
 6    no compound-related signs of gross pathology nor histopathologic lesions indicative of
 7    noncancer toxicological effects in response to methanol; however, they did not provide any
 8    detailed data to illustrate these findings.
 9           As discussed above and in Section 5.1.1, reproductive and developmental effects are
10    considered the most sensitive and quantifiable effects reported in studies of methanol.  Oral
11    reproductive and developmental studies employed single doses that were too high to be of use.
12    In the absence of suitable reproductive or developmental data from oral exposure studies, it was
13    decided to conduct a route-to-route extrapolation and to use the critical effect from the inhalation
14    study (brain weight) to derive an RfD. Thus, the POD (in terms of AUC methanol in blood) used
15    for the derivation of the RfC was also used for the derivation of the RfD.  This POD was divided
16    by a UF of 100, and a human PBPK model was used to obtain an RfD value of 0.4 mg/kg-day.
17    As for the RfC, the 100-fold UF includes a UFn of 10 for intraspecies variability, a UFA of 3 to
18    address pharmacodynamic uncertainty, and a UFD of 3 for database uncertainty.
19           The confidence in the RfD is medium to high. Despite the relatively high confidence in
20    the critical studies, all limitations to confidence as presented for the RfC also  apply to the RfD.
21    Confidence in the RfD is slightly lower than for the RfC due to the lack of adequate oral studies
22    for the RfD derivation, necessitating a route-to-route extrapolation.

      6.2.3. Cancer/Oral and Inhalation
23           Under the current Guidelines for  Carcinogen Risk Assessment (U.S. EPA 2005a, 2005b),
24    methanol fulfils the criteria to be described as likely to be a human carcinogen by all routes of
25    exposure. This descriptor is based principally on findings of dose-related, statistically significant
26    increases in the incidence of: lymphoreticular tumors in lifetime studies of both sexes of Eppley
27    Swiss Webster mice (Apaja, 1980) and both sexes of Sprague-Dawley rats (Soffritt et al., 2002a),
28    a slight but significant (compared to historical controls) increase in relatively  rare hepatocellular
29    carcinomas in male Sprague-Dawley rats following oral  exposure (Soffritti et al., 2002a), and
30    dose-related occurrences of pulmonary adenomas/adenocarcinomas and pheochromocytomas in
31    F344 rats by  inhalation exposure (NEDO, 1985/2008b).  This determination is supported by the
32    results of other studies that have shown tumorigenic responses similar to those observed by
33    Soffritti et al. (2002a) in rats exposed to formaldehyde, a metabolite of methanol, and to the
34    metabolic precursors of methanol and formaldehyde, aspartame and MTBE. In addition,
35    epidemiological studies have associated exposure to formaldehyde with increases in the
                                                6-6      DRAFT—DO NOT CITE OR QUOTE

-------
 1   incidence of both leukemias and lymphomas (IARC, 2004). However, the key studies, Soffritti
 2   et al. (2002a), NEDO (1985/2008b) and Apaja (1980), have associated uncertainties (see below
 3   and discussions in Sections 4.9.2 and 5.4.3) that reduce confidence in the chosen descriptor.
 4          The statistically significant increase in the incidence of lymphoreticular tumors observed
 5   in the Soffritti et al. (2002a) drinking water study of Sprague-Dawley rats was used in the
 6   determination of the POD for estimating the methanol oral CSF. APBPK model was developed,
 7   and several model predictions of internal dose metrics were considered for use in the dose-
 8   response analysis and derivation of the human equivalent dose. Methanol metabolized was
 9   selected as the dose metric best suited for derivation of the oral POD because of its superior fit to
10   the response data and consistency with the hypothesis that formaldehyde may be the
11   carcinogenic agent associated with methanol exposure. The EPA multistage cancer model was
12   used to derive a BMDLio for the male rat in terms of mg methanol metabolized/day. Assuming
13   that metabolized methanol distributes in the body according to body weight to the 3/4 power, the
14   rat BMDLio of 63.9 mg-day was converted to a human BMDLio of 3,553 mg-day.  The human
15   PBPK model was then used to convert this human mg-day value for total methanol metabolized
16   back to a human equivalent methanol oral dose HED(BMDLio) of 51.5 mg/kg-day for
17   lymphomas in the male rat.  The oral CSF of 2E-03  (mg/kg-day)"1 was then derived based on a
18   linear extrapolation from this POD to estimated background levels.
19          Pulmonary adenomas/adenocarcinomas in male F344 rats and pheochromocytomas in
20   female F344 rats observed in the chronic inhalation  study of NEDO (1985/2008b) were
21   considered in the determination of the POD for estimating the inhalation CSF.  In this case, all
22   dose metrics estimated by the PBPK model provided a similar acceptable fit to the tumor
23   response data.  Methanol metabolized was  selected as  the dose metric for derivation of the
24   inhalation POD for consistency with the approach used for the derivation of the oral POD and
25   with the hypothesis that formaldehyde may be the carcinogenic agent associated with methanol
26   exposure. As for the oral POD, the EPA multistage cancer model was used to derive a BMDLio
27   for the rat in terms of mg methanol metabolized/day. Assuming that metabolized methanol
28   distributes in the body according to body weight to the % power, the rat BMDLio of 15 mg-day
29   was converted to a human BMDLio of 971 mg-day.  The human PBPK model was then used to
30   convert this human mg-day value  for total methanol metabolized back to a human equivalent
31   methanol inhalation concentration F£EC(BMCLio) of 81,900 ug/m3 for pheochromocytomas in
32   the female rat. The inhalation cancer unit risk of 1E-06 (ug/m3)"1 was then derived based on a
33   linear extrapolation from this POD to estimated background levels.
34          Section 5.4.3 of this assessment documents several uncertainties with the quantification
35   of cancer risk.  The main uncertainties can be grouped into issues related to study quality, the
36   interpretation of study results, and the consistency of the results with other laboratories.  Other
37   uncertainties discussed in Section 5.4.3 include the choice of tumor endpoint, the choice of dose-

                                               6-7       DRAFT—DO NOT CITE OR QUOTE

-------
1   response model, the PBPK model and dose metric used for the animal to human extrapolations,
2   and the human relevance of the carcinogenic responses in rats and mice.
                                            6-8      DRAFT—DO NOT CITE OR QUOTE

-------
                                        7. REFERENCES

 1          Abbondandolo, A; Bonatti, S; Corsi, C; et al. (1980) The use of organic solvents in
 2   mutagenicity testing. Mutat Res 79:141-150.
 O
 4          ACGIH (American Conference of Governmental Industrial Hygienists). (2000) Threshold
 5   Limit Values for Chemical Substances and Physical Agents and Biological Exposure Indices.
 6   American Conference of Governmental Industrial Hygienists. Cincinnati, OH.
 7
 8          Adanir, J; Ozkalkanti, MY; Aksun, M. (2005) Percutaneous methanol intoxication: Case
 9   report. Eur J Anaesthiol 22:560-561.
10
11          Agarwal, DP. (2001) Genetic polymorphisms of alcohol metabolizing enzymes. Pathol
12   Biol (Paris) 49(9):703-709.
13
14          Airas, L; Paavilainen, T; Marttila, RJ; et al.  (2008) Methanol intoxication-induced
15   nigrostriatal dysfunction detected using 6-[18F]fluoro-L-dopa PET. Neurotoxicol 29:671-674.
16
17          Akaike, H. (1973) Information theory and an extension of the maximum likelihood
18   principle. Petrov, BN; Csaki, F; eds. Second International Symposium on Information Theory,
19   September 1971, Tsahkadsor, Armenia, USSR. Budapest, Hungary: Akademia Kiado.
20   pp.267-281.
21
22          Albert RE, Sellakumar AR, Laskin S, Kuschner M, Nelson N, Snyder CA.  1982.
23   Gaseous formaldehyde and hydrogen chloride induction of nasal cancer in the rat.  J Natl Cancer
24   Inst. 68(4):597-603.
25
26          Albin, RL and Greenamyre, JT. (1992) Alternative excitotoxic hypotheses. Neurology
27   42:733-738.
28
29          Allen, BC; Kavlock, RJ; Kimmel, CA; et al. (1994) Dose-response assessment for
30   developmental toxicity II: comparison of generic benchmark dose estimates with no observed
31   adverse effect levels. Fundam Appl Toxicol 23:487-495.
32
33          Andresen, H; Schmoldt, H; Matschke, J; et  al. (2008) Fatal methanol intoxication with
34   different survival times -morphological findings and postmortem methanol distribution. Forensic
35   Sci Intl 179:206-210.
36

                                              7-1         DRAFT - DO NOT CITE OR QUOTE

-------
 1          Andrews, JE; Ebron-McCoy, M; Logsdon, TR; et al. (1993) Developmental toxicity of
 2   methanol in whole embryo culture: a comparative study with mouse and rat embryos. Toxicology
 3   81(3):205-215.
 4
 5          Andrews, JE; Ebron-McCoy, M; Kavlock, RJ; et al. (1995) Developmental toxicity of
 6   formate and formic acid in whole embryo culture: a comparative study with mouse and rat
 7   embryos. Teratology 51:243-251.
 8
 9          Andrews, JE; Ebron-McCoy, M; Schmid, JE; et al. (1998) Effects of combinations of
10   methanol and formic acid on rat embryos in culture. Teratology 58(2):54-61.
11
12          Andrews, LS; Clary, JJ; Terrill, JB; et al. (1987) Subchronic inhalation toxicity of
13   methanol. J Toxicol  Environ Health 20:117-124.
14
15          Ang, HL; Deltour, L; Hayamizu, TF; et al. (1996) Retinoic acid synthesis in mouse
16   embryos during gastrulation and craniofacial development linked to class IV alcohol
17   dehydrogenase gene expression. J Biol Chem 271:9526-9534.
18
19          Antony, AC. (2007) In utero physiology: role of folic acid in nutrient delivery and fetal
20   development. Am J Clin Nutr 85(suppl):598S- 603S.
21
22          Apaja, M.  (1980) Evaluation of toxicity and carcinogenicity of malonaldehyde. Acta
23   Universitatis Ouluensis Series D Medica 55 Anatomica, Pathologica, Microbiologica No. 8, 1-
24   59.
25
26          Arora, V; Nijjar, IS; Thukral, H; et al. (2005) Bilateral putaminal necrosis caused by
27   methanol intoxication- a case report. Ind J Radiol Imag 15:341-342.
28
29          Arora, V; Nijjar, IBS; Multani, AS; et al. (2007) MRI findings in methanol intoxication: a
30   report of two cases. Brit J Radiol 80:e243-e246.
31
32          Aschner, M; Kimmelberg, HK. (1996) The Role of Glia in Neurotoxicity, Published by
33   CRC Press, p.365.
34
35          ATSDR (Agency for Toxic Substances and Disease Registry). (1997) Toxicological
36   review of methyl tertiary butyl ether. Agency for Toxic  Substances and Disease Registry,
37   Department of Health, Atlanta, GA.
38
                                               7-2         DRAFT - DO NOT CITE OR QUOTE

-------
 1          ATSDR (Agency for Toxic Substances and Disease Registry). (1999) Toxicological
 2   review of formaldehyde. Agency for Toxic Substances and Disease Registry, Department of
 3   Health, Atlanta, GA.
 4
 5          Aufderheide, TP; White, SM; Brady, WJ; et al. (1993) Inhalational and percutaneous
 6   methanol toxicity in two firefighters. Ann Emergency Med 22(12): 1916-1918.
 7
 8          Aziz, MH; Agrawal, AK; Adhami, VM; et al. (2002) Methanol-induced neurotoxicity in
 9   pups exposed during lactation through mother: role of folic acid. Neurotoxicol Teratol 24(4):
10   519-527.
11
12          Azmak, D. (2006) Methanol related deaths in Edirne. Legal Med 8:35-42.
13
14          Barceloux, DG; Bond, GR; Krenzelok, EP; et al. (2002) American Academy of Clinical
15   Toxicology practice guidelines on the treatment of methanol poisoning. J Toxicol Clin Toxicol
16   40(4):415-446.
17
18          Battelle. (1981) Final report on a chronic inhalation toxicology study in rats and mice
19   exposed to formaldehyde. Prepared for the Chemical Industry Institute of Toxicology. Prepared
20   by Battelle Columbus Laboratories, Columbus, OH. CUT Docket # 10922.
21
22          Batterman, SA; Franzblau, A. (1997) Time-resolved cutaneous absorption and permeation
23   rates of methanol in human volunteers. Int Arch Occup Environ Health 70(5):341-351.
24
25          Batterman, SA; Franzblau, A; D'Arcy, JB; et al. (1998) Breath, urine, and blood
26   measurements as biological exposure indices of short-term inhalation exposure to methanol.
27   Int Arch Occup Environ Health 71:325-335.
28
29          Beam, P; Patel, J; O'Flynn, WR.  (1993) Cervical ribs: a cause of distal and cerebral
30   embolism. Postgrad Med J 69:65-68.
31
32          Bebarta, VS; Heard, K; Dart, RC. (2006) Inhalational abuse of methanol products:
33   elevated methanol and formate levels without vision loss. Am J Emerg Med 24:725-728.
34
35          Belpoggi F; Soffritti M; Guarino  M; et al. (2002a) Results of long-term experimental
36   studies on the carcinogenicity of Ethylene-bis-Dithiocarbamate (Mancozeb) in rats. Ann NY
37   AcadSci 982:123-136.
38
                                               7-3         DRAFT - DO NOT CITE OR QUOTE

-------
 1          Belpoggi F; Soffritti M; Minardi F; et al. (2002b) Results of long-term carcinogenicity
 2   bioassays on tert-amyl-methyl-ether (TAME) and di-isopropyl-ether (DIPE) in rats.  Ann N Y
 3   Acad Sci 982:70-86.
 4
 5          Belpoggi, F; Soffritti, M; Maltoni, C. (1995) Methyl-tertiary-butyl ether (MTBE) - a
 6   gasoline additive - causes testicular and lymphohaematopoietic cancers in rats. Toxicol Ind
 7   Health 11:119-149.
 8
 9          Belpoggi, F; Soffritti, M; Filippini, F; et al. (1997) Results of long-term experimental
10   studies on the carcinogenicity of methyl tert-butyl ether. Ann NY Acad Sci 837:77-95.
11
12          Belpoggi, F; Soffritti, M; Maltoni, C. (1998) Pathological characterization of testicular
13   tumours and lymphomas-leukaemias, and of their precursors observed in Sprague-Dawley rats
14   exposed to methy-tertiary-butyl-ether (MTBE). Eur J Oncol 3(3):201-206.
15
16          Belpoggi, F; Soffritti, M; Maltoni, C. (1999) Immunoblastic lymphomas in Sprague-
17   Dawley rats following exposure to the gasoline oxygenated additives Methyl-Tertiary-Butyl
18   Ether (MTBE) and Ethyl-Tertiary-Butyl Ether (ETBE): early observations on their natural
19   history. Eur J Oncol 4:563-572.
20
21          Bennett, IL, Jr; Gary, FH; Mitchell, GL, Jr; et al. (1953) Acute methyl alcohol poisoning:
22   a review based on experiences in an outbreak of 323 cases. Medicine 32:431-463.
23
24          Benton, CD; Calhoun, FP Jr. (1952) The ocular effects of methyl alcohol poisoning:
25   report of a catastrophe involving three hundred and twenty persons. Trans Am Acad Ophthalmol
26   Otolaryngol 56(6):875-885.
27
28          Bessell-Browne, RJ; Bynevelt, M. (2007) Two cases of methanol poisoning:  CT and MRI
29   features. Aust Radiol 51:175-178.
30
31          Bhatia, KP; Marsden, CD. (1994) The behavioral and motor consequences of focal
32   lesions of the basal ganglia in man. Brain 117(Pt 4): 859-876.
33
34          Black, KA; Eells,  JT; Noker, PE; et al. (1985) Role of hepatic tetrahydrofolate in the
35   species difference in methanol toxicity. Proc  Natl Acad Sci 82:3854-3858.
36
                                               7-4         DRAFT - DO NOT CITE OR QUOTE

-------
 1          Blancato, JN; Evans, MV; Power, FW; et al. (2007) Development and use of PBPK
 2   modeling and the impact of metabolism on variability in dose metrics for the risk assessment of
 3   methyl tertiary butyl ether (MTBE). J Environ Protect Sci  1:29-51.
 4
 5          Blanco, M; Casado, R; Vasquez; et al. (2006) CT and MR imaging findings in methanol
 6   intoxication. Am J Neuroradiol 27:452-454.
 7
 8          Bosron, WF; Li, TK. (1986) Genetic polymorphism of human liver alcohol and aldehyde
 9   dehydrogenases, and their relationship to alcohol metabolism and alcoholism. Hepatology
10   6(3):502-510.
11
12          Bouchard, M; Brunet, RC; Droz, PO; et al. (2001) A biologically based dynamic model
13   for predicting the disposition of methanol and its metabolites in animals and humans. Toxicol Sci
14   64:169-184.
15
16          Boutwell, RK. (1964) Some biological aspects of skin carcinogenesis. Prog Exp Tumor
17   Res 4:207-250.
18
19          Braden GL; Strayhorn CH;  Germain MJ;  et al. (1993) Increased osmolal gap in alcoholic
20   acidosis. Arch Intern Med 153(20):2377-80.
21
22          Brahmi, N; Blel, Y; Abidi, N; et al. (2007) Methanol poisoning in Tunisia:  report of 16
23   cases. Clin Toxicol 45(6):717-20.
24
25          Branch, S; Rogers, JM; Brownie, CF; et al. (1996) Supernumerary lumbar rib:
26   manifestation of basic alterations in embryonic development of ribs. J. Appl. Toxicol.
27   16:115-119.
28
29          Brent, J; Lucas, M; Kulig, K; et al. (1991) Methanol poisoning in a six-week-old infant.
30   JPediatr 118:644-646.
31
32          Brien, JF; Clarke, DW; Richardson, B; et al. (1985) Disposition of ethanol in maternal
33   blood, fetal blood, and amniotic fluid of third-trimester pregnant ewes. Am J Obstet Gynecol
34   1985 Jul 1;152(5):583-90.
35
36          Brown, RP; Delp, MD; Lindstedt, SL; et al. (1997) Physiological parameters values for
37   physiologically based pharmacokinetic models. Toxicol Ind Health 13:407-484.
38

                                               7-5         DRAFT - DO NOT CITE OR QUOTE

-------
 1          Bucher, JR. (2002) The National Toxicology Program rodent bioassay: designs,
 2    interpretations, and scientific contributions. Ann N Y Acad Sci. Dec; 982:198-207.
 3
 4          Burbacher, TM. (1993) Neurotoxic effects of gasoline and gasoline consituents.  Environ
 5    Health Perspect 101:133-141.
 6
 7          Burbacher, T; Shen, D; Grant, K; et al. (1999a) Methanol disposition and reproductive
 8    toxicity in adult females. Reproductive and offspring developmental effects following maternal
 9    inhalation exposure to methanol in nonhuman primates. HEI Research Report Number 89: Part I:
10    9-68.
11
12          Burbacher, T; Grant, K; Shen, D; et al. (1999b) Developmental effects in infants exposed
13    prenatally to methanol. Reproductive and offspring developmental effects following maternal
14    inhalation exposure to methanol in nonhuman primates. HEI Research Report Number 89: Part I:
15    69-117.
16
17          Burbacher, TM; Grant, KS; Shen, DD; et al. (2004a) Chronic maternal methanol
18    inhalation in nonhuman primates (M. fascicularis): reproductive performance and birth outcome.
19    Neurotoxicol Teratol 26:639-650.
20
21          Burbacher, TM; Shen, DD; Lalovic, B; et al. (2004b) Chronic methanol inhalation in
22    nonhuman primates {M. fascicularis}:  exposure and toxicokinetics prior to and during pregnancy.
23    Neurotoxicol Teratol 26:201-221.
24
25          Burnell, JC; Li, T-K; Bosron, WE (1989) Purification and steady-state kinetic
26    characterization of human liver b3b3 alcohol dehydrogenase. Biochemistry 28:6810-6815.
27
28          Burwell RD, Whealin J, Gallagher M. (1992) Effects of aging on the diurnal pattern of
29    water intake in rats. Behav Neural Biol. Nov;58(3): 196-203. PubMed PMID: 1456941.
30
31          Butchko, HH; Stargel, WW; Comer, CP; et al. (2002) Aspartame: review of safety. Regul
32    Toxicol Pharmacol 35:81-93.
33
34          Calabrese, EJ. (2001) Assessing the default assumption that children are always at risk.
35    Hum Ecol Risk Assess 7(l):37-59.
36
37          Caldwell, J; Jinot, J; Devoney,  D; et al. (2008) Evaluation of evidence for infection as a
38    MOAfor induction of rat lymphoma. Environ Mol Mutagen 49(2): 155-164.
                                               7-6         DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2          Caldwell, J; Jinot, J; Devoney, D; et al. (2009) Author reply to comment on "Evaluation
 3    of evidence for infection as the mode of action for induction of rat lymphoma." Environ Mol
 4    Mutagen. Jan;50(l): 6-9.
 5
 6          Cameron, AM; Nilsen, OG; Haug, E; et al. (1984) Circulating concentrations of
 7    testosterone, luteinizing hormone and follicle stimulating hormone in male rats after inhalation
 8    of methanol. Arch Toxicol 7(Suppl):441-443.
 9
10          Cameron, AM; Zahlsen, K; Haug, E; et al. (1985) Circulating steroids in male rats
11    following inhalation of n-alcohols. Arch Toxicol 8(Suppl):422-424.
12
13          Campbell, JA; Howard, DR; Backer, LC; et al. (1991) Evidence that methanol inhalation
14    does not induce chromosome damage in mice. Mutat Res 260:257-264.
15
16          Carson, BL; McCann, JL; Ells, JV III; et al. (1981) Methanol Health Effects. Prepared by
17    the Midwest Research Institute. Prepared for the Office of Mobile Source Air Pollution Control,
18    U. S. Environmental Protection Agency, Ann Arbor, MI. PB82-160797.
19
20          Caspi, R; Foerster, H; Fulcher, CA; et al. (2006) MetaCyc: a multiorganism database of
21    metabolic pathways and enzymes. Nucleic Acids Res 34:D511-6.
22
23          CERHR (Center for the Evaluation of Risks to Human Reproduction). (2004)
24    NTP-CERHR Expert Panel report on the reproductive and developmental toxicity of methanol.
25    Reprod Toxicol 18:303-390.
26
27          Chemfinder. (2002) Methanol. Available at http://chemfmder.cambridgesoft.com/.
28
29          Chen, JC; Schneiderman, JF; Wortzman, G. (1991) Methanol poisoning: bilateral
30    putaminal and  cerebellar cortical lesions on CT and MR.  J Comput Assist Tomogr 15(3):
31    522-524.
32
33          Chen-Tsi, C.  (1959) Materials on the hygienic standardization of the maximally
34    permissible concentration of methanol vapors in the atmosphere. Gig Sanit 24:7-12 (as cited in
35    Carson, et al.,  1981).
36
                                               7-7         DRAFT - DO NOT CITE OR QUOTE

-------
 1          Chernoff, N; Roger, JM. (2004) Supernumerary ribs in developmental toxicity bioassays
 2    and in human populations: incidence and biological significance. J Toxicol Environ Health
 3    7:437_449.
 4
 5          Chuwers, P; Osterloh, J; Kelly, T; et al. (1995) Neurobehavioral effects of low-level
 6    methanol vapor exposure in healthy human volunteers. Environ Res 71(2): 141-150.
 7
 8          Cichoz-Lach, H; Partycka, J; Nesina, I; et al. (2007) Genetic polymorphism of alcohol
 9    dehydrogenase 3 in digestive tract alcohol damage.  Hepatogastroenterology 54(76): 1222-7.
10
11          Clarke, DW; Steenaart, NA; Brien,  F. (1986) Disposition of ethanol and activity of
12    hepatic and placental alcohol dehydrogenase and aldehyde dehydrogenases in the third-trimester
13    pregnant guinea pig for single and short-term oral ethanol administration. Alcohol Clin Exp Res
14    10(3):330-6.
15
16          Clancy, B; Finlay, BL; Darlington, RB; et al. (2007) Extrapolating brain development
17    from experimental species to humans. Neurotoxicology 28(5):931-7.
18
19          Clary, JJ. (2003) Methanol, is it a developmental risk to humans? Regul Toxicol
20    Pharmacol 37:83-91.
21
22          Clewell, HJ III; Gentry, PR; Gearhart, JM; et al. (2001) Development of a physiologically
23    based pharmacokinetic model of isopropanol and its metabolite acetone. Toxicol Sci 63:160-172.
24
25          Cogliano, VJ; Grosse, Y; Baan, RA; et al. (2005) Meeting report: summary of IAEC
26    monographs on formaldehyde, 2-butoxyethanol, and l-tert-butoxy-2-propanol. Environ Health
27    Perspect 113:1205-1208.
28
29          Coleman, CN; Mason, T; Hooker, ER; et al.  (1999) Developmental effects of intermittent
30    prenatal exposure to 1,1,1-trichloroethane in the rat. Neurotoxicol Teratol 21(6):699-708.
31
32          Connell, JL; Doyle, JC; Gurry, JF. (1980) The vascular complications of cervical ribs.
33    AustNZJSurg 50:125-130.
34
35          Cook, MR; Bergman, FJ; Cohen, HD; et al. (1991) Effects of methanol vapor on human
36    neurobehavioral measures. HEI Research Report Number 42.
37

                                               7-8         DRAFT - DO NOT CITE OR QUOTE

-------
 1          Cook, RJ; Champion, KM; Giometti, CS. (2001) Methanol toxicity and formate oxidation
 2   in NEUT2 mice. Arch Biochem Biophys 393:192-198.
 3
 4          Cooper, RL; Mole, ML; Rehnberg, GL; et al. (1992) Effect of inhaled methanol on
 5   pituitary and testicular hormones in chamber acclimated and nonacclimated rats. Toxicology
 6   71:69-81.
 7
 8          Corley, RA; Barmett, GA; Ghanayem, BI. (1994) Physiologically based
 9   pharmacokinetics of 2-butoxyethanol and its major metabolite, 2-butoxyacetic acid, in rats and
10   humans. Toxicol Appl Pharmacol 129:61-79.
11
12          Corley, RA; Mast, TJ; Carney, EW; et al. (2003) Evaluation of physiologically based
13   models of pregnancy and lactation for their application in children's health risk assessments. Crit
14   Rev Toxicol 33:137-211.
15
16          Crebelli, R; Conti, G; Conti, L; et al. (1989) A comparative study on ethanol and
17   acetaldehyde  as inducers of chromosome malsegregation in Aspergillus nidulans. Mutat Res
18   215:187-195.
19
20          Cronholm, T. (1987) Effect of ethanol on the redox state of the coenzyme bound to
21   alcohol dehydrogenase studied in isolated hepatocytes. Biochem J 248:567-572.
22
23          Crump, KS. (1984) A new method for determining allowable daily intakes. Fundam Appl
24   Toxicol 4:854-871.
25
26          Crump, KS. (1995) Calculation of benchmark doses from continuous data. Risk Anal
27   15:79-89.
28
29          Cruzan, G. (2009) Assessment of the cancer potential of methanol. Critical Reviews in
30   Toxicology, 2009; 39(4): 347-363.
31
32          Cruzan, G; Cushman, JR; Andrews, LS; Granville, GC; Johnson, KA; Hardy, CJ;
33   Coombes, DW; Mullins, PA; Brown, WR. (1998) Chronic toxicity/oncogenicity study of styrene
34   in CD rats by inhalation exposure for 104 weeks. Toxicol Sci 46:266-281.
35
36          Cumming, ME; Ong, BY; Wade, JG; et al. (1984) Maternal and fetal ethanol
37   pharmacokinetics and cardiovascular responses in near-term pregnant sheep. Can J Physiol
38   Pharmacol 62(12): 1435-9.
                                              7-9         DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2          Davis, VE; Brown, H; Huff, JA; et al. (1967) The alteration of serotonin metabolism to
 3    5-hydroxytrypophol by ethanol ingestion in man. J Lab Clin Med 69:132-140.
 4
 5          Davoli, E; Cappellini, L; Airoldi, L; et al. (1986) Serum methanol concentrations in rats
 6    and in men after a single dose of aspartame. Food Chem Toxicol 24:187-189.
 7
 8          D'Alessandro, A; Osterloh, JD; Chuwers; et al. (1994) Formate in serum and urine after
 9    controlled methanol exposure at the threshold limit value. Environ Health Perspect 102:178.
10
11          De Brabander, N; Wojciechowski, M; De Decker, K; et al. (2005) Fomepizole as a
12    therapeutic strategy  in paediatric methanol poisoning. Eur J Pediatr 164:158-161.
13
14          De Flora, S;  Zanacchi, P; Camoirano, A; et al. (1984) Genotoxic  activity and potency of
15    135 compounds in the Ames reversion test and in a bacterial DNA-repair test. Mutat Res
16    133:161-198.
17
18          Degitz, SJ; Zucker, RM; Kawanishi, CY; et al. (2004a) Pathogenesis of methanol-induced
19    craniofacial defects  in C57BL/6J mice. Birth Defects Res. A 70:172-178.
20
21          Degitz, SJ; Rogers, JM; Zucker, RM; et al.  (2004b) Developmental toxicity of methanol:
22    pathogenesis in CD-I and C57BL/6J mice exposed in whole embryo culture. Birth Defects Res
23    A70:179-184.
24
25          Deltour, L; Foglio, MH; Duester, G. (1999) Metabolic deficiencies in alcohol
26    dehydrogenase ADH1, ADH3, and ADH4 null mutant. Overlapping roles of ADH1 and ADH4 in
27    ethanol clearance and metabolism of retinol to retinoic acid. J Biol Chem 274:16796-16801.
28
29          Dethlefs, R;  Naraqi, S. (1978) Ocular manifestations and complications of acute methyl
30    alcohol intoxication. Med J Aust 2(10):483-485.
31
32          Devore, JL. (1995) Probability and statistics for engineering and  the sciences (4th ed).
33    Belmont, CA: Duxbury.
34
35          Dicker, E; Cedebaum, AL (1986) Inhibition of the low Km mitachondrial aldehyde
36    dehydrogenase by diethyl maleate and phorone in vivo and in vitro implications for
37    formaldehyde metabolism. Biochem J 240:821-828.
38
                                              7-10        DRAFT - DO NOT CITE OR QUOTE

-------
 1          Dikalova, AE; Kadiiska, MB; Mason, RP (2001) An in vivo ESR spin-trapping study:
 2   free radical generation in rats from formate intoxication-role of the Fenton reaction. Proc Natl
 3   AcadSci 98:13549-13553.
 4
 5          Dorman, DC; Moss, OR; Farris, GM; et al. (1994) Pharmacokinetics of inhaled (14C)
 6   methanol and methanol-derived 14-C formate in normal and folate-deficient cynomolgus
 7   monkeys. Toxicol Appl Pharmacol 128(2):229-238.
 8
 9          Dorman, DC; Bolon, B; Struve, MF; et al. (1995) Role of formate in methanol-induced
10   exencephaly in CD-I mice. Teratology 52(1):30-40.
11
12          Downie, A; Khattab, TM; Malik, MI; et al. (1992) A case of percutaneous industrial
13   methanol toxicity. Occup Med (London) 42(1): 47-49.
14
15          Dudka, J. (2006) The total antioxidant status in the brain after ethanol or 4-methyl-
16   pyrazole administration to rats intoxicated with methanol. Exp Toxicol Pathol 57:445-448.
17
18          Dutkiewicz, B; Konczalik, J; Karwacki, W. (1980) Skin absorption and per os
19   administration of methanol in men. Int Arch Occup Environ Health 47(l):81-88.
20
21          Eells, JT; Black, KA; Tedford, CE; et al. (1983) Methanol toxicity in the monkey: effects
22   of nitrous oxide and methionine. J Pharmacol Exp Ther 227:349-353.
23
24          EFSA (European Food Safety Authority). (2006) Opinion of the scientific panel on food
25   additives, flavourings, processing aids and materials in contact with food (AFC) on a request
26   from the commission related to a new long-term carcinogenicity study on aspartame. The EFSA
27   Journal 356:1-44.
28
29          EFSA (European Food Safety Authority). (2009) Updated Scientific Opinion of the Panel
30   on Food Additives and Nutrient Sources added to Food on a request from the European
31   Commission related to the 2nd ERF carcinogenicity study on aspartame, taking into consideration
32   study data submitted by the Ramazzini Foundation in February 2009.  The EFSA Journal
33    1015:1-18.
34
35          Ernstgard, L; Shibata, E;  Johanson, G. (2005) Uptake and disposition of inhaled methanol
36   vapor in humans. Toxicol Sci 88:30-38.
37

                                              7-11         DRAFT - DO NOT CITE OR QUOTE

-------
 1          Ernstgard, L. (2005) Personal communication. E-mail correspondance from Lena
 2   Ernstgard to Torka Poet dated March 23, 2005 [see Section B.3.5 of Appendix B].
 3
 4          Estonius, M; Svensson, S; H00g, JO. (1996) Alcohol dehydrogenase in human tissues:
 5   localization of transcripts coding for five classes of the enzyme. FEES Lett 397:338-342.
 6
 7          Evans, AL. (1999) Pseudoseizures as a complication of painful cervical ribs. Dev Med
 8   Child Neurol 41:840-842.
 9
10          Everitt JI, Richter CB. (1990) Infectious diseases of the upper respiratory tract:
11   implication for toxicology studies. Environ Health Perspect 85:239-247.
12
13          Pagan, JF; Singer, LT. (1983) Infant recognition memory as a measure of intelligence. In:
14   Advances in Infancy Research, Ed, Lipsitt, LP Ablex, New York, NY (as cited in Burbacher
15   etal., 1999a)
16
17          Fallang, B; Saugstad, OD; Gragaard, J; et al. (2003) Kinematic quality of reaching
18   movements in preterm infants. Pediatr Res 53:836-842.
19
20          Feany, MB; Anthony, DC; Frosch, MP; et al. (2001) August 2000: two cases with
21   necrosis and hemorrhage in the putamen and white matter. Brain Pathol 11(1): 121-122.
22
23          Fernandez Noda, El; Nunez-Arguelles, J; Perez Fernandez, J; et al. (1996) Neck and
24   brain transitory vascular compression causing neurological complications, results of surgical
25   treatment on  1300 patients. J Cardiovasc Surg 37:155-166.
26
27          Finkelstein, Y; Vardi, J. (2002) Progressive Parkinsonism in a young, experimental
28   physicist following long-term exposure to methanol. Neurotoxicol 23:521-525.
29
30          Fiserova-Bergerova, V; Diaz, ML. (1986) Determination and prediction of tissue-gas
31   partition coefficients. Int Arch Occup Environ Health 58:75-87.
32
33          Fisher, JW; Dorman, DC; Medinsky, MA; et al. (2000) Analysis of respiratory exchange
34   of methanol in the lung of the monkey using a physiological  model. Toxicol Sci 53:185-193.
35
36          Fontenot, AP; Pelak, VS. (2002) Development of neurologic symptoms in a 26-year-old
37   woman following recovery from methanol intoxication. Chest 122(4): 1436-1438.
38
                                              7-12         DRAFT - DO NOT CITE OR QUOTE

-------
 1          Foster, MW; Stamler, J. (2004) New insights into protein s-nitrosylation-
 2    mitochondria as a model system. 279(24):25891-25897.
 3
 4          Frederick, LJ; Schulte, PA; Apol, A. (1984) Investigation and control of occupational
 5    hazards associated with the use of spirit duplicators. Am Ind Hyg Assoc J 45:51-55.
 6
 7          Fu, SS; Sakanashi, TM; Rogers, JM; et al. (1996) Influence of dietary folic acid on the
 8    developmental toxicity of methanol and the frequency of chromosomal breakage in the CD-I
 9    mouse. Reprod Toxicol 10:455-463.
10
11          Gaiter, D; Carmine, A; Buervenich, S; et al. (2003) Distribution of class I, III, and IV
12    alcohol dehydrogenase mRNAs in the adult rat, mouse and human brain. Eur J Biochem
13    270:1316-1326.
14
15          Gannon KS, Smith JC, Henderson R, Hendrick P. (1992) A system for studying the
16    microstructure of ingestive behavior in mice. Physiol Behav. Mar;51(3):515-21. PubMed PMID:
17    1523228.
18
19          Gaul, HP; Wallace, CJ; Auer, RN et al. (1995) MR findings in methanol intoxication. Am
20    JNeuroradiol  16:1783-1786.
21
22          Gentry, PR; Covington, TR; Andersen, ME; et al. (2002) Application of a physiologically
23    based pharmacokinetic model for isopropanol in the derivation of a reference dose and reference
24    concentration. Regul Toxicol Pharmacol 36:51-68.
25
26          Gentry, PR; Covington, TR; Clewell, HJ III. (2003) Evaluation of the potential impact of
27    pharmacokinetic differences on tissue dosimetry in offspring during pregnancy and lactation.
28    Regul Toxicol Pharmacol 38:1-16.
29
30          Gibson, MAS; Butters, NS; Reynolds, JN; et al. (2000) Effects of chronic prenatal
31    ethanol exposure on locomotor activity, and hippocampal weight, neurons, and nitric oxide
32    synthase activity of the young postnatal guinea pig. Neurotoxicol Teratol 22:183-192.
33
34          Gonzalez-Quevedo, A; Obregon, F; Urbina, M; et al. (2002) Effect of chronic methanol
35    administration on amino acids and monoamines in retina, optic nerve, and brain of the rat.
36    Toxicol Appl Pharmacol 185(2):77-84.
37

                                               7-13         DRAFT - DO NOT CITE OR QUOTE

-------
 1          Griffiths, AJF. (1981) Neurospora and environmentally induced aneuploidy. In Stich, HF;
 2    San, RHC. eds. Short-term Tests for Chemical Carcinogens. Springer-Verlag, pp. 187-199.
 3
 4          Guerri, C and Sanchis, R. (1985) Acetaldehyde and alcohol levels in pregnant rats and
 5    their fetuses. Alcohol 2(2):267-70.
 6
 7          Guggenheim, MA; Couch, JR; Weinberg, W. (1971) Motor dysfunction as a permanent
 8    complication of methanol ingestion. Arch Neurol 24:550-554.
 9
10          Hailey, JR. (2004) Chairperson's report: lifetime study in rats conducted by the
11    Ramazzini Foundation. Prepared by J.R. Hailey (pathology working group chair) NIEHS and
12    submitted to F. Belpoggi, November 30, 2004.
13
14          Hansen, JM; Contreras, KM; Harris, C. (2005) Methanol, formaldehyde, and sodium
15    formate exposure in rat and mouse conceptuses: a potential role of the visceral yolk sac in
16    embryotoxicity. Birth Defects Res (Part A)  73:72-82.
17
18          Hantson, P-E. (2005) Acute methanol intoxication: physiopathology, prognosis and
19    treatment. Bull Mem Acad R Med Belg 160:294-300.
20
21          Hantson, P; Duprez,  T; Mahieu, P. (1997a) Neurotoxicity to the basal ganglia shown by
22    magnetic resonance imaging (MRI) following poisoning by methanol and other substances.
23    J Toxicol Clin Toxicol  35(2): 151-161.
24
25          Hantson, P; Lambermont, JY; Mahieu, P. (1997b) Methanol poisoning during late
26    pregnancy. J Toxicol Clin Toxicol  35(2): 187-191.
27
28          Hanzlik, RP; Fowler, SC; Eells, JT.  (2005) Absorption and elimination of formate
29    following oral administration of calcium formate in female human subjects. Drug Metab Dispos
30    33:282-286.
31
32          Hassanian-Moghaddam, H; Pajoumand, A; Dadgar, SM; et al.. (2007) Prognostic factors
33    in methanol poisoning. Human Exp Toxicol 26(7):583-6.
34
35          Harris,  C; Wang, S-W; Lauchu, JJ; et al.  (2003) Methanol metabolism and
36    embryotoxicity in rat and mouse conceptuses: comparison of alcohol dehydrogenase (ADH1),
37    formaldehyde dehydrogenase (ADH3), and CAT. Reprod Toxicol 17:349-357.
38

                                              7-14        DRAFT - DO NOT CITE OR QUOTE

-------
 1          Harris, C; Dixon, M; Hansen, JM. (2004) Glutathione depletion modulates methanol,
 2   formaldehyde, and formate toxicity in cultured rat conceptuses. Cell Biol Toxicol 20:133-145.
 3
 4          Harris, N; Jaffe, E; Diebold, J; et al. (1999) World Health Organization classification of
 5   neoplastic diseases of the hematopoietic and lymphoid tissues: report of the clinical advisory
 6   committee meeting-Airlie house, Virginia, November 1997. J Clin Oncol 17(12):3835-3849.
 7
 8          Haseman JK, Hailey JR, Morris RW (1998) Spontaneous neoplasm incidences in Fischer
 9   344 rats and B6C3F1 mice in two-year carcinogenicity studies: ANational Toxicology Program
10   update. Toxicol Pathol 26:428-441.
11
12          Hass, U; Lund, SP; Simonsen, L; et al. (1995) Effects of prenatal exposure to xylene on
13   postnatal development and behavior in rats. Neurotoxicol Teratol  17(3):341-349.
14
15          Hauptmann, M; Lubin, JH; Stewart, PA; et  al. (2003) Mortality from
16   lymphohematopoietic malignancies among workers in formaldehyde industries. J Natl Cancer
17   Inst 95:1615-1623.
18
19          Health Review Committee. (1999) Commentary.  HEI Research Report Number 89:
20   pp.119-133.
21
22          Heck, H; Casanova, M. (2004) The implausibility of leukemia induction by
23   formaldehyde: a critical review of the biological evidence on distant-site toxicity. Regul Toxicol
24   Pharmacol 40:92-106.
25
26          Hedberg, JJ; Backlund, M; Stromberg, P; et al. (2001) Functional polymorphisms in the
27   alcohol dehydrogenase 3 (ADH3) promoter. Pharmacogenetics ll(9):815-824.
28
29          HEI. (1987) Special Report: Automotive methanol vapors and human health: An
30   evaluation of existing scientific information and issues for further research.
31
32          Heidelberger, C; Freeman, AE; Pienta, RJ; et al. (1983) Cell transformation by chemical
33   agents — a review and analysis of the literature. Areport of the U.S. Environmental Protection
34   Agency Gene-Tox Program. Mutat Res 114:283-385.
35
36          Henderson, MS. (1914) Cervical rib. Report of thirty-one cases. Am J Orthop Surg
37   11:408-430.
38
                                               7-15         DRAFT - DO NOT CITE OR QUOTE

-------
 1          Henderson, WR; Brubacher, J. (2002) Methanol and ethylene glycol poisoning: a case
 2    study and review of current literature. Can J Emerg Med 4:34-40.
 3
 4          Hess, DT; Matsumoto, A; Kim, SO; et al. (2005) Protein S-nitrosylation: purview and
 5    parameters. Nat Rev Mol Cell Biol 6:150-166
 6
 7          Hinderer, RK; Johnson, MN, Seppala, A; Jacobs, MM; Apaja, M; Patil, KD; Shubik, P.
 8    (1979) Toxicity and carcinogenicity of Agerite Resin D (ARD), trimethyldihydroquinoline
 9    polymer. Submitted for publication to J. Env.  Path. Toxicol. (as cited in Apaja, 1980)
10
11          Hines, RN; McCarver, DG. (2002) The ontogeny of human drug-metabolizing enzymes:
12    phase I oxidative enzymes. J Pharmacol Exp Ther 300:355-360.
13
14          Horton, VL; Higuchi,  MA; Rickert, DE. (1992) Physiologically based pharmacokinetic
15    model for methanol in rats, monkeys, and humans. Toxicol Appl Pharmacol 117:26-36.
16
17          Hovda, KE; Hunderi,  OH; Tafjord, A-B; et al. (2005) Methanol outbreak in Norway
18    2002-2004; epidemiology, clinical features and prognostic signs. J Int Med 258:181-190.
19
20          Hovda, KE; Mundal, H; Urdal, P; et al. (2007) Extremely slow formate elimination in
21    severe methanol poisoning: a fatal case report. Clin Toxicol 45:516-521.
22
23          HSDB  (Hazardous Substances Data Bank). (2002) Available at
24    http://toxnet.nlm.nih.gov/cgi-bin/sis/htmlgen7HSDB.
25
26          Hsu, HH; Chen, CY; Chen, FH et al. (1997) Optic atrophy and cerebral infarcts caused by
27    methanol intoxication. Neuroradiol 39:192-194.
28
29          Huang, QF; Gebrewold, A; Zhang, A; et al. (1994) Role of excitatory amino acids in
30    regulation of rat pial microvasculature. Am J Physiol 266:R158-R163.
31
32          Huff, J. (2002) Chemicals studied and evaluated in long-term carcinogenesis bioassays
33    by both the Ramazzini Foundation and the National Toxicology Program: In tribute to Cesare
34    Maltoni and David Rail. Ann  NY Acad Sci  982208-982230.
35
36          Huff, J; LaDou, J. (2007) Aspartame bioassay findings portend human cancer hazards. Int
37    J Occup Environ HealthOct-Dec;13(4):446-8.
38
                                              7-16        DRAFT - DO NOT CITE OR QUOTE

-------
 1          Huff, J; Jacobson, MF and Davis, DL. (2009) The Limits of Two-Year Bioassay Exposure
 2   Regimens for Identifying Chemical Carcinogens. Environ Health Perspect 116: 1439-1442.
 3
 4          Hunderi, OH; Hovda, KE; Jacobsen, D. (2006) Use of the osmolal gap to guide the start
 5   and duration of dialysis in methanol poisoning. Scand J Urol Nephrol 40:70-74.
 6
 7          IARC (International Agency for Research on Cancer). (2004) Formaldehyde,
 8   2-butoxyethanol and l-tert-butoxy-2-propanol.  Summary of data reported and evaluation. IARC
 9   Monographs on the Evaluation of Carcinogenic Risks to Humans 88:1-7.
10
11          Im, H;  Mun, J; Khim, JY; et al. (2006) Evaluation of toxicological monitoring markers
12   using proteomic analysis in rats exposed to formaldehyde. J Proteome Res 5:1354-1366.
13
14          Infurna, R; Weiss, B. (1986) Neonatal behavioral toxicity in rats following prenatal
15   exposure to methanol. Teratology 33:259-265.
16
17          Ingemansson, SO. (1984) Clinical observations on ten cases of methanol poisoning with
18   respect to ocular manifestations. Acta Opthalmolog (Copenh) 62(1): 15-24.
19
20          IPCS (International Programme on Chemical Safety). (1997) Environmental Health
21   Criteria 196 - Methanol 180.
22
23          Ishii, H. (1981) Incidence of brain tumors in rats fed aspartame. Toxicol Lett 7:433^137
24   (as cited in EFSA, 2006).
25
26          Ishii, H; Koshimizu, T; Usami, S; et al. (1981) Toxicity of aspartame and its
27   diketopiperazine for Wistar rats by dietary administration for 104 weeks. Toxicol Lett 21:91-94
28   (as cited in EFSA, 2006)
29
30          Jacobsen, D; McMartin, KE. (1986) Methanol and ethylene glycol poisonings
31   Mechanism of toxicity, clinical course, diagnosis and treatment. Med Toxicol 1:309-334.
32
33          Jacobsen, D; Webb, R; Collins, TD; et al. (1988) Methanol  and formate kinetics in late
34   diagnosed methanol intoxication. Med Toxicol 3:418-423.
35
36          Johanson,  G; Kronborg, H; Naslund, PH; et al. (1986) Toxicokinetics of inhaled
37    2 butoxyethanol (ethylene glycol monobutyl ether) in man. Scand  J Work Environ Health 12(6):
38   594-602.
                                              7-17        DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2          Johlin, FC; Fortman, CS; Nghiem, DD; et al. (1987) Studies on the role of folic acid and
 3   folate-dependent enzymes in human methanol poisoning. Mol Pharmacol 31(5):557-61.
 4
 5          Kahn, A; Blum, D (1979) Methyl alcohol poisoning in an 8-month-old boy: an unusual
 6   route of intoxication. J Pediatrics 94(5): 841-843.
 7
 8          Kamata E, Nakadate M, Uchida O, Ogawa Y, Suzuki S, Kaneko T, Saito M, Kurokawa Y.
 9   1997. Results of a 28-month chronic inhalation toxicity study of formaldehyde in male Fisher-
10   344 rats. J Toxicol Sci. 22(3):239-54.
11
12          Kavet, R; Nauss, K. (1990) The toxicity of inhaled methanol vapors. Crit Rev Toxicol
13   21(1):21-50.
14
15          Kavlock, RJ; Allen, BC; Faustman, EM; et al. (1995) Dose-response assessments for
16   developmental toxicity. IV: Benchmark doses for fetal weight changes. Fund Appl Toxicol 26:
17   211-222.
18
19          Kawai, T; Yasugi, T; Mizunma, K; et al. (1991) Methanol in urine as a biological
20   indicator of occupational exposure to methanol vapor. Int Arch Occup Environ Health
21   63:311-318.
22
23          Keles, GT; Orgii9, S; Toprak, B; et al. (2007) Methanol poisoning with necrosis corpus
24   callosum. Clin Toxicol 45:307-308.
25
26          Kerns, WD; Pavkov, KL; Donofrio, DJ; et al. (1983) Carcinogenicity of formaldehyde in
27   rats and mice after long-term inhalation exposure. Cancer Res 43:4382-4392.
28
29          Kerns, WII; Tomaszewski, C; McMartin, K; et al. META Study Group. (2002) Formate
30   kinetics in methanol poisoning. J Toxicol Clin Toxicol 40:137-143.
31
32          Keys DA, Schultz IR, Mahle DA, Fisher JW. (2004) A quantitative description of suicide
33   inhibition of dichloroacetic acid in rats and mice. Toxicol Sci. Dec;82(2):381-93. Sep 16.
34   PubMedPMID: 15375292.
35
36          Kim, SW; Jang, YJ; Chang, JW; et al. (2003) Degeneration of the nigrostriatal pathway
37   and induction of motor deficit by tetrahydrobiopterin: an in vivo model  relevant to Parkinson's
38   disease. Neurobiology of Disease 13:167-176.

                                              7-18        DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2          Knowles, K. (2008) Personal communication. E-mail correspondance from Kathryn
 3   Knowles to Jeff Gift dated June 14, 2008.
 4
 5          Kraut, JA; Kurtz, I. (2008) Toxic alcohol ingestions: clinical features, diagnosis, and
 6   management. Clin J Am Soc Nephrol 3(l):208-225.
 7
 8          Kuteifan, K; Oesterle, H; Tajahmady, T; et al., (1998) Necrosis and hemorrhage of the
 9   putamen in methanol poisoning shown on MRI. Neuroradiol 40:158-160.
10
11          Leakey JEA, Seng JE, Allaben WT. 2003. Body weight considerations in the B6C3F1
12   mouse and the use of dietary control to standardize background tumor incidence in chronic
13   bioassays. Toxicol Appl Pharmacol 193:237-265.
14
15          Lee, EW; Brady, AN; Brabec, MJ; et al. (1991) Effects of methanol vapors on
16   testosterone production and testis morphology in rats. Toxicol Ind Health 7(4):261-275.
17
18          Lee, EW; Terzo, TS; D'Arcy, JB; et al. (1992) Lack of blood formate accumulation in
19   humans following exposure to methanol vapor at the current permissible exposure limit of
20   200 ppm. Am IndHyg Assoc J 53:99-104.
21
22          Lee, EW; Garner, CD; Terzo, TS. (1994) Animal model for the study of methanol
23   toxicity: comparison of folate-reduced rat responses with published monkey data. J Toxicol
24   Environ Health 41:71-82.
25
26          Lewis, RJ Jr. (1992) Sax's dangerous properties of industrial materials. 8th edition Van
27   Nostrand Reinhold, New York, NY.
28
29          Ley, CO; Gali, FG (1983) Parkinsonian syndrome after methanol intoxication. Eur
30   Neurol 22:405-409.
31
32          Lilly, MZ; Gold, KW; Rogers, RR. (1995)  Integrating QA prinicples with basic elements
33   of a research program promotes quality  science in a nonGLP research laboratory. In vitro Toxicol
34   8:191-198.
35
36          Lindsey, JR; Davidson, MK; Schoeb, TR; Cassell, GH. (1985)Mycoplasmapulmonis-
37   host relationships in a breeding colony of Sprague-Dawley rats with enzootic murine respiratory
38   mycoplasmosis. Lab Anim Sci 35:597-608.
                                              7-19         DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2          Liu, JJ; Daya, MR; Carrasquillo, O; et al. (1998). Prognostic factors in patients with
 3    methanol poisoning. J Toxicol Clin Toxicol 36(3):175-181.
 4
 5          Liu, L; Hausladen, A; Zeng, M; et al. (2001) A metabolic enzyme for S-nitrosothiol
 6    conserved from bacteria to humans. Nature 410:490-494.
 7
 8          Lorente, C; Cordier, S; Bergeret, A; et al. (2000) Maternal occupational risk factors for
 9    oral clefts. Scand J Work Environ Health 26(2): 137-145.
10
11          Mahieu, P; Hassoun, A; Lauwerys, R; et al. (1989) Predictors of methanol intoxication
12    with unfavorable outcome. Hum Toxicol 8(2): 135-137.
13
14          Makar, AB; Tephly,  TR; Mannering, GJ. (1968) Methanol metabolism in the monkey.
15    MolPharmacol 4:471-483.
16
17          Maltoni, C; Lefemine, G; Cotti, G (1986) Experimental research on trichlorothylene
18    carcinogenesis. Princeton: Princeton Scientific Publishing Co.
19
20          Maltoni C, Ciliberti A, Cotti G, Perino G. 1988. Long-term carcinogenicity bioassays on
21    acrylonitrile administered by inhalation and by  ingestion to Sprague-Dawley rats. Ann N Y Acad
22    Sci. 534:179-202.
23
24          Maltoni, C; Ciliberti, A; Pinto, C; et al. (1997) Results of long-term experimental
25    carcinogenicity studies of the effects of gasoline, correlated fuels, and major gasoline aromatics
26    on rats. Ann NY Acad Sci 837:15-52.
27
28          Mann, WJ; Muttray, A; Schaefer, D; et al. (2002) Exposure to 200 ppm of methanol
29    increases the concentrations of interleukin-lbeta and interleukin-8 in nasal secretions of healthy
30    volunteers.  Ann Otol Rhinol Laryngol 111:633-638.
31
32          Mannering, GJ; Van Harken, DR; Makar, AB; et al. (1969) Role of the intracellular
33    distribution of hepatic CAT in the peroxidative  oxidation of methanol. Ann NY Acad  Sci
34    168:265-280.
35
36          Maronpot, RR; Flake, G; Huff, J. (2004) Relevance of animal carcinogenesis findings to
37    human cancer predictions and prevention. Toxicol Pathol. Mar-Apr;32 Suppl 1:40-8.  Review.
38
                                               7-20         DRAFT - DO NOT CITE OR QUOTE

-------
 1          McGregor, DB; Martin, R; Riach, CG; et al. (1985) Optimization of a metabolic
 2    activation system for use in the mouse lymphoma L5178Y tk+tk" system. Environ Mutagen
 3    7(Suppl3):10.
 4
 5          McConnell, E; Solleveld, H; Swenberg, J; et al. (1986) Guidelines for combining
 6    neoplasms for evaluation of rodent carcinogenesis studies. J Natl Cancer Inst 76(2):283-289.
 7
 8          McKellar, MJ; Hidajat, RR; Elder, MJ. (1997) Acute ocular toxicity: clinical and
 9    electrophysiological features Aust NZ J Opthalmol. 25(3):225-230.
10
11          Medinsky, MA; Dorman, D; Bond, J; et al. (1997) Pharmacokinetics of methanol and
12    formate in female cynomolgus monkeys exposed to methanol vapors. HEI Research Report
13    Number 77:1-38.
14
15          Melnick RL, Thayer KA, and Bucker JR.  2007. Conflicting views on chemical
16    carcinogenesis arising from the design and evaluation of rodent carcinogenicity studies. Environ
17    Health Perspect 116:130-135; doi:10.1289/ehp.9989 (availableatdx.doi.org).
18
19          Meister, A; Anderson, ME. (1983) Glutathione.  Ann Rev Biochem 52:711-760.
20
21          Methanol Institute. (2006) Methanol Institute Milestones. Washington, D.C. Available at
22    http ://www.methanol. org
23
24          Meyer, RJ; Beard, ME; Ardagh, MW; et al. (2000) Methanol poisoning. NZ Med J
25    113(1102):11-13.
26
27          Molinary,  S.V. 1984. Preclinical studies of aspartame in nonprimate animals. In
28    Aspartame: Physiology and Biochemistry (L.D. Stegink and LJ. Filer, Jr., Eds.), pp. 289-306.
29    Marcel Dekker, Inc.,  New York.
30
31          Molotkov, A;  Deltour, L; Foglio, MH; et al.  (2002) Distinct retinoid metabolic functions
32    for alcohol dehydrogenase genes ADH1 and ADH4 in protection against vitamin A toxicity or
33    deficiency revealed in double null mutant mice. J Biol Chem 277:13804-13811.
34
35          Montserrat, CA; Field, MS; Perry, C; Ghandour, H; Chiang, E; Selhub, J; Shane, B; and
36    Stover, PJ. (2006) Regulation of Folate-mediated One-carbon Metabolism by 10-
37    Formyltetrahydrofolate Dehydrogenase. The Journal of Biological Chemistry 281(27): 18335-
38    18342.
                                              7-21         DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2          Mooney, SM; Miller, MW. (2001) Episodic exposure to ethanol during development
 3    differentially effects brainstem nuclei in the macaque. J Neurocytol 30:973-982.
 4
 5          Muthuvel, A; Rajamani, R; Senthilvelan, M; et al. (2006a) Modification of allergenicity
 6    and immunogenicity of formate dehydrogenase by conjugation with linear mono methoxy poly
 7    ethylene glycol: improvement in detoxification of formate in methanol poisoning. Clin Chim
 8    Acta 374:122-128.
 9
10          Muthuvel, A; Rajamani, R; Manikandan, S; et al. (2006b) Detoxification of formate by
11    formate dehydrogenase-loaded erythrocytes and carbicarb in folate-deficient methanol-
12    intoxicated rats. Clin Chim Acta 367:162-169.
13
14          Nagasawa, H; Wada, M; Koyama, S; et al. (2005) A case of methanol intoxication with
15    optic neuropathy visualized on STIR sequence of MR images. Clin Neurol 45:527-530.
16
17          Naraqi, S; Dethlefs,  RF; Slobodniuk, RA; et al. (1979) An outbreak of acute methyl
18    alcohol intoxication Aust N Z  J Med 9(l):65-68.
19
20          NCI (National Cancer  Institute). (2006) Prospective study of aspartame-containing
21    beverages and risk of hematopoietic and brain cancers. Abstract No. 4010. Available from the
22    American Association for Cancer Research (97th AACR Annual Meeting, Washington, DC) at
23    http://www.aacr.org (as cited in EFSA,  2006).
24
25          NEDO (New Energy Development Organization). (1987) Toxicological research of
26    methanol as a fuel for power station. Summary report on tests with monkeys,  rats and mice. New
27    Energy Development Organization, Tokyo, Japan. Unpublished report.
28
29          NEDO. (1985/2008a) Test Report:  18-month inhalation carcinogenicity study on
30    methanol in B6C3F1 mice (Test No. 4A-223). Eight volume translation (consisting of report
31    text, summary tables, individual  animal tables and copies of pathology slides) of original report
32    prepared by the Mitsubishi Kasei Institute  of Toxicology and Environmental Sciences (currently
33    Mitsubishi  Chemical Safety Insitute, Ltd.) for the New Energy Development Organization,
34    Tokyo, Japan. (Original report prepared in 1985; Translation submitted to the EPA IRIS hotline
35    by the Methanol Insitute on March 28, 2008 and certified as accurate and complete by NEDO in
36    letter dated May 12, 2008). Unpublished report.
37

                                               7-22         DRAFT - DO NOT CITE OR QUOTE

-------
 1          NEDO. (1985/2008b) Test Report: 24-month inhalation carcinogenicity study on
 2    methanol in Fischer rats (Test No. 5A-268).  Eight volume translation (consisting of report text,
 3    summary tables, individual animal tables and copies of pathology slides) of original report
 4    prepared by the Mitsubishi Kasei Institute of Toxicology and Environmental Sciences (currently
 5    Mitsubishi Chemical Safety Insitute, Ltd.) for the New Energy Development Organization,
 6    Tokyo, Japan. (Original report prepared in 1985; Translation submitted to the EPA IRIS hotline
 7    by the Methanol Insitute on March 28, 2008 and certified as accurate and complete by NEDO in
 8    letter dated May 12, 2008). Unpublished report.
 9
10          Nelson, BK; Brightwell, WS; MacKenzie, DR; et al. (1985) Teratological assessment of
11    methanol and ethanol at high inhalation levels in rats.  Fund Appl Toxicol 5:727-736.
12
13          Nelson, J. (1967) Respiratory infections of rats and mice with emphasis on indigenous
14    mycoplasms. In: Pathology of laboratory rats and mice. Cotchin, E; Rose, F, eds. Blackwell
15    Scientific Publications, Oxford and Edinburgh, G.B. pp. 259-294.
16
17          Nguyen, T; Baumgartner, F; Nelems, B. (1997) Bilateral rudimentary first ribs as a cause
18    of thoracic outlet syndrome. J Natl Med Assoc 89:69-73.
19
20          Nilsson, JA; Hedberg, JJ; Vondracek, M; et al. (2004) Alcohol dehydrogenase 3
21    transcription associates with proliferation of human oral keratinocytes. Cell Mol Life Sci
22    61:610-617.
23
24          Noker, PE;  Eells, JT; Tephly, TR; et al. (1980) Methanol toxicity: treatment with folic
25    acid and 5-formyl tetrahydrofolic acid. Alcohol Clin Exp Res 4:378-383.
26
27          NRC (National Research Council).  (1994) Science and judgment in risk assessment.
28    Prepared by the Committee on Risk Assessment of Hazardous Air Pollutants, Board on
29    Environmental Studies and Toxicology, Commission on Life Sciences. National Academy Press,
30    Washington, D.C.
31
32          NRC (National Research Council).  (1983) Risk assessment in the federal government:
33    managing the process. Washington, DC: National Academy Press.
34
35          NTP (National Toxicology Program). (1990) Toxicology and Carcinogenesis  Studies of
36    Toluene (CAS No.  108-88-3) in F344/N Rats  and B6C3F1 Mice (Inhalation Studies). Public
37    Health Service, U.S. Department of Health and Human Services; NTP TR 371; Available from

                                              7-23        DRAFT - DO NOT CITE OR QUOTE

-------
 1   the National Institute of Environmental Health Sciences, Research Triangle Park, NC and online
 2   at http://ntp.mehs.nih.gov/ntp/htdocs/LT_rpts/tr371 .pdf.
 3
 4          NTP (National Toxicology Program). (1999) Toxicology Data Management System,
 5   Historical Control Tumor Incidence Summary. - Inhalation Studies of Fischer 344 Rats. Report
 6   updated 12/99. Available at http://ntp.niehs.nih.gov.
 7
 8          NTP (National Toxicology Program). (2000) NTP technical report on the toxicology and
 9   carcinogenesis studies of 2 butoxyethanol (CAS No.  Ill 76 2) in F344/N rats and B6C3F1 mice
10   (inhalation studies). Public Health Service, U.S. Department of Health and Human Services;
11   NTP TR 484; NIH Publ. No. 00-3974. Available from the National Institute of Environmental
12   Health Sciences, Research Triangle Park, NC and online at
13   http://ntp.niehs.nih.gov/ntp/htdocs/LT_rpts/tr484.pdf.
14
15          NTP (National Toxicology Program). (2003) NTP Technical Report. Toxicity studies of
16   aspartame in FVB/N-TgN(v-Ha-ras)Led (Tg.AC) hemizygous mice and carcinogenicity studies
17   of aspartame in B6.129-Trp53m&Brd (N5) haploinsufficient mice. NTP GMM 1, Research
18   Triangle Park, NC (as cited in EFSA, 2006).
19
20          NTP (National Toxicology Program). (2006) Specifications for the conduct of studies to
21   evaluate the toxic and carcinogenic potential of chemical, biological  and physical agents in
22   laboratory animals for the national toxicology program (NTP). Available at
23   http://ntp.Niehs.Nih.Gov/files/specifications 2006octl.Pdf
24
25          NTP (National Toxicology Program). (2007) NTP Historical Controls Report, All Routes
26   and Vehicles, Rats. Report prepared October, 2007. Available at http://ntp.niehs.nih.gov.
27
28          OECD (Organization for Economic Co-operation and Development ). (2007)
29   Establishment and control of archives that  operate in compliance with the principles  of GLP
30   OECD Environment, Health and Safety Publications. No. 15. OECD Environment Directorate,
31   Paris, France.
32
33          Ohno, K; Tanaka-Azuma,  Y; Yoneda, Y; et al. (2005) Genotoxicity test system based on
34   p53R2 gene expression in human  cells: Examination with 80 chemicals. Mutat Res 588:47-57.
35
36          Onder, F; Ilker, S; Kansu,  T; et al. (1999) Acute blindness and putaminal necrosis in
37   methanol intoxication. Int Opthalmol 22:81-84.
38
                                               7-24         DRAFT - DO NOT CITE OR QUOTE

-------
 1          Osterloh, JD; D'Alessandro, A; Chuwers, P; et al. (1996) Serum concentrations of
 2    methanol after inhalation at 200 ppm. J Occup Environ Med 38:571-576.
 3
 4          Pace, V; Perentes, E; Germann, PG. (2002) Pheochromocytomas and ganglioneuromas in
 5    the aging rats: morphological and immunohistochemical characterization.  Toxicol Pathol
 6    30(4):492-500.
 7
 8          Parthasarathy, NJ; Kumar, RS; Karthikeyan, P; et al. (2005a) In vitro and in vivo study of
 9    neutrophil functions after acute methanol intoxication in albino rats. Toxicol Environ Chem
10    87:559-566.
11
12          Parthasarathy, NJ; Kumar, RS; Devi, RS. (2005b) Effect of methanol intoxication on rat
13    neutrophil functions. J Immunotoxicol 2:115-211.
14
15          Parthasarathy, NJ; Kumar, RS; Manikandan, S; et al. (2006a) Effect of methanol-induced
16    oxidative stress on the neuroimmune system of experimental rats. Chemico-Biological Inter
17    161:14-25.
18
19          Parthasarathy, NJ; Kumar, RS; Manikandan, S; et al. (2006b) Methanol-induced oxidative
20    stress in rat lymphoid organs. J Occup Health 48:20-27.
21
22          Parthasarathy, NJ; Kumar, RS; Manikandan, S; et al. (2007) Effect of methanol
23    intoxication on specific immune functions of albino rats. Cell Biol Toxicol 23:177-187.
24
25          Pastino, GM; Conolly, RB. (2000) Application of a physiologically based pharmaco-
26    kinetic model to estimate the bioavailability of ethanol in male rats: distinction between gastric
27    and hepatic pathways of metabolic clearance. Toxicol Sci 55(2):25665.
28
29          Patankar, T; Bichile, L; Karnad, D; et al. (1999) Methanol poisoning: computed
30    tomography scan findings in four patients. Austral Radiol 43(4):526-528.
31
32          Pelletier, JU; Habib, MH; Khalil, R; et al.  (1992) Putaminal necrosis after methanol
33    intoxication. J Neurol Neurosurg Psychiatry 55(3): 234-235.
34
35          Peng, MT; Chen, YT; Hung, SH; Yaung, CL. (1990) Circadian rhythms of feeding and
36    drinking behavior of rats aged from 3 to 21 months. Proc Natl Sci Counc Repub China B.
37    Apr;14(2):98-104. PubMed PMID: 2247537.
38

                                              7-25        DRAFT - DO NOT CITE OR QUOTE

-------
 1          Perkins, RA; Ward, KW; Pollack, GM. (1995a) Comparative toxicokinetics of inhaled
 2   methanol in the female CD-I mouse and Sprague-Dawley rat. Fundam Appl Toxicol 28:
 3   245-254.
 4
 5          Perkins, RA; Ward, KW; Pollack, GM. (1995b) A pharmacokinetic model of inhaled
 6   methanol in humans and comparison to methanol disposition in mice and rats. Environ Health
 7   Perspect 103:726-73 3.
 8
 9          Perkins, RA; Ward, KW; Pollack, GM. (1996) Methanol inhalation: site and other factors
10   influencing absorption, and an inhalation toxicokinetic model for the rat. Pharm Res 13:749-755.
11
12          Phang, PT; Passerini, L; Mielke, B; et al. (1988) Brain hemorrhage associated with
13   methanol poisoning. Crit Care Med 16:137-140.
14
15          Pietruszko, R. (1980) Alcohol and aldehyde dehydrogenase isozymes from mammalian
16   liver-their structural and functional differences. Isozymes Curr Top Biol Med Res 4:107-130.
17
18          Pikkarainen, PH; Raiha, NCR. (1967) Development of alcohol dehydrogenase activity in
19   the human liver. Pediatr Res 1(3): 165-168.
20
21          Pinkerton, LE; Hein, MJ; Stayner,  LT. (2004) Mortality among a cohort of garment
22   workers exposed to formaldehyde: an update. Occup Environ Med 61:193-200.
23
24          Pitkin, RM. (2007) Folate and neural tube defects. Am J Clin Nutr 85:2858-2888.
25
26          Plantinga, Y; Perdock, J; de Groot, L.  (1997) Hand function in low-risk preterm infants:
27   its relation to muscle power  regulation. Dev Med Child Neurol 39:6-11.
28
29          Poet, TS; Teeguarden, JG; Hinderliter, PM. (2006) Final Report: Development,
30   calibration and application of a methanol PBPK model. Prepared for the National Center for
31   Environmental Assessment,  U.S. EPA. Prepared by the Center for Biological Monitoring and
32   Modeling, Battelle Pacific Northwest National Laboratory, Richland, WA.
33
34          Pollack, GM; Brouwer, KL. (1996) HEI Research Report Number 74: Maternal-Fetal
35   Pharmacokinetics of Methanol. 58.
36
37          Pollack, GM; Brouwer, KL; Kawagoe, JL. (1993) Toxicokinetics of intravenous methanol
38   in the female rat. Fundam Appl Toxicol 21:105-110.
                                              7-26         DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2          Pollack, GM; Kawagoe, JL. (1991) Determination of methanol in whole blood by
 3    capillary gas chromatography with direct on-column injection. J Chromatogr B Biomed Appl
 4    570:406-11.
 5
 6          Poon, R; Chu, IH; Bjarnason, S; et al. (1994) Inhalation toxicity study of methanol,
 7    toluene and methanol/toluene mixtures in rats: effects of 28-day exposure. Toxicol Ind Health
 8    10(3):231-245.
 9
10          Poon, R; Chu, IH; Bjarnason, S; et al. (1995) Short-term inhalation toxicity of methanol,
11    gasoline and methanol/gasoline in the rat. Toxicol Ind Health 11(3):343-361.
12
13          Prabhakaran, V; Ettler, H; Mills, A; et al. (1993) Methanol poisoning: two cases with
14    similar plasma methanol concentrations but different outcomes. Can Med Assoc J 148(6):
15    981-984.
16
17          Que, LG; Yan, Y; Whitehead, GS; et al. (2005) Protection from experimental asthma by
18    an endogenous bronchodilator. Science 308:1618-1621.
19
20          Rajamani, R; Muthuvel, A; Senthilvelan, M; et al. (2006) Oxidative stress induced by
21    methotrexate alone and in the presence of methanol in discrete regions of the rodent brain, retina
22    and optic nerve.  Toxicol Lett 165:265-273.
23
24          Ramazzini Foundation. 2006. European Ramazzini Foundation stands behind aspartame
25    study results, announces ongoing research on artificial sweeteners. Available at
26    http://www.ramzzini/it/eng/fondazione/eventidettagli.asp?id5292
27
28          Ramsey,  JC; Andersen, ME. (1984) A physiologically based description of the inhaled
29    pharmacokinetics of styrene in rats and humans. Toxicol Appl Pharmacol 73:159-175.
30
31          Rappaport, H. (1966) Tumors of the hematopoietic system. In: Atlas of Tumor Pathology,
32    Sect III. Fasc. 9: Armed Forces Institute of Pathology. Wahsington, D.C., pp. 49-64 [as cited in
33    Apaja, 1989]
34
35          Razin, S; Yogev, D; Naot, Y. (1998) Molecular biology and pathogenicity of
36    mycoplasmas. Microbiol Mol Biol Rev. Dec; 62(4): 1094-156. Review.
37

                                               7-27        DRAFT - DO NOT CITE OR QUOTE

-------
 1          Rice, D; Barone, S Jr. (2000) Critical periods of vulnerability for the developing nervous
 2    system: evidence from humans and animal models. Environ Health Perspect 108(Suppl 3):
 3    511-33.
 4
 5          Riegel, H; Wolf, G. (1966) Severe neurological deficiencies as a consequence of methyl
 6    alcohol poisoning. Fortschr Neurol Psychiat 34:346-351.
 7
 8          Rogers, JM; Barbee, BD; Mole, ML. (1995) Exposure concentration and time (CxT)
 9    relationships in the developmental toxicity of methanol in mice. Toxicologist 15:164.
10
11          Rogers, JM; Barbee, BD; Rhenberg, BF. (1993b) Critical periods of sensitivity  for the
12    developmental toxicity of inhaled methanol. Teratology 47:395.
13
14          Rogers, JM; Brannen, KC; Barbee, BD; et al. (2004) Methanol exposure during
15    gastrulation causes holoprosencephaly, facial dysgenesis, and cervical vertebral malformations in
16    C57BL/6J mice.  Birth Defects Res B 71:80-88.
17
18          Rogers, JM; Mole, ML. (1997) Critical periods of sensitivity to the developmental
19    toxicity of inhaled methanol in the CD-I mouse. Teratology 55(6):364-372.
20
21          Rogers, JM; Mole, ML; Chernoff, N; et al. (1993a) The developmental toxicity of inhaled
22    methanol in the CD-I mouse, with quantitative dose-response modeling for estimation of
23    benchmark doses. Teratology 47(3): 175-88.
24
25          Rogers, VV; Wickstrom, M; Liber, K; et al. (2002) Acute and Subchronic Mammalian
26    Toxicity  of Naphthalenic Acids from Oil Sands Tailings.  Toxicol Sci 66:347-355.
27
28          Rotenstreich, Y; Assia, El; Kesler, A. (1997) Late treatment of methanol blindness. Brit J
29    Opthalmol 81:415-420.
30
31          Rubinstein, D; Escott, E; Kelly, JP (1995) Methanol intoxication with putaminal and
32    white matter necrosis: MR and CT findings. Am J Neuroradiol 16(7): 1492-1494.
33
34          Sakanashi, TM; Rogers, JM; Fu,  SS; et al. (1996) Influence of maternal folate status on
35    the developmental toxicity of methanol in the CD-I mouse. Teratology 54:198-206.
36
37          Salzman, M. (2006) Methanol  neurotoxicity. Clin Toxicol 44:89-90.
38

                                               7-28        DRAFT - DO NOT CITE OR QUOTE

-------
 1          Saxton, EH; Miller, TQ; Collins, JD. (1999) Migraine complicated by brachial
 2   plexopathy as displayed by MRI and MRA: aberrant subclavian artery and cervical ribs. J Natl
 3   MedAssoc 91:333-341.
 4
 5          Sayers, RR; Yant, WP; Schrenk, HH; et al. (1944) Methanol poisoning: II. Exposure of
 6   dogs for brief periods eight times daily to high concentrations of high methanol vapor in air.
 7   J. Ind Hyg Toxicol 28:255-259.
 8
 9          Schaefer, PW; Grant, PE; Gonzalez, RG (2000) Diffusion-weighted MR imaging of the
10   brain. Radiol 217(2):331-345.
11
12          Schoeb, TR; Mcconnell, EE; Juliana, MM; Davis, JK; Davidson, MK; and Lindsey, JR.
13   (2009) Mycoplasmapulmonis and Lymphoma in Bioassays in Rats. Vet Pathol 46.
14
15          Schulte, PA; Burnett, CA; Boeniger, MF; et al. (1996) Neurodegenerative diseases:
16   occupational occurrence and potential risk factors,  1982-1991. Am J Public Health 86:
17   1281-1288.
18
19          Schumacher, R; Mai, A; Gutjahr, P.  (1992) Association of rib abnormalities and
20   malignancy in childhood. Eur J Pediatr  151:432-434.
21
22          Scrimgeour, EM; Dethlefs, RF; Kevau, I. (1982) Delayed recovery of vision after
23   blindness caused by methanol poisoning. Med J Aust 2(10):481-483.
24
25          Sedivec, V; Mraz, M;  Flek, J. (1981) Biological monitoring of persons exposed to
26   methanol vapors. Int Arch Occup Environ Health 48:257-271.
27
28          Sefidbakht, S; Rasekhi, AR; Kamali, K; et al. (2007) Methanol poisoning:  acute MR and
29   CT findings in nine patients. Neuroradiol 49:427-435.
30
31          Sellakumar AR, Snyder CA,  Solomon JJ, Albert RE. 1985. Carcinogenicity of
32   formaldehyde and hydrogen chloride in rats. Toxicol Appl Pharmacol. 81(3 Pt l):401-6.
33
34          Short, DW. 1975. The subclavian artery in 16 patients with complete cervical ribs.
35   J Cardiovasc Surg  16:135-141.
36
                                              7-29         DRAFT - DO NOT CITE OR QUOTE

-------
 1          Simintzi, I; Schulpis, KH; Angelogianni, P; et al. (2007) The effect of aspartame
 2    metabolites on the suckling rat frontal cortex acetylcholinesterase. An in vitro study. Food Chem
 3    Toxicol 45:2397-2401.
 4
 5          Simmon, VF; Kauhanen, K; Tardiff, RG. (1977) Mutagenic activity of chemicals
 6    identified in drinking water. Scott, D; Bridges, B; Sobel, F. eds. Progress in Genetic Toxicology.
 7    Vol 2. Elsevier/North Holland Press; pp. 249-268.
 8
 9          Sinkeldam, E; Kuper, C; Beems, R; et al. (1991) Combined chronic toxicity and
10    carcinogenicity study with acesulfame-k in rats. Acesulfame-k (Mayer, D; Kemper, F, eds.). New
11    York: Marcel Dekker; pp. 43-58.
12
13          Siragusa, RJ; Cerda, JJ; Baig, MM; et al. (1988) Methanol production from the
14    degradation of pectin by human colonic bacteria. Am J Clin Nutr. 47:848-51
15
16          Skrzydlewska, E; Elas, M; Ostrowska, J. (2005) Protective effects of N-acetylcysteine
17    and vitamin E derivative U83836E on proteins modifications induced by methanol intoxication.
18    Toxicol Mechanisms Methods 15:263-270.
19
20          Smith, ME; Newman, HW. (1959) The rate of ethanol metabolism in fed and fasting
21    animals. J Biol Chem 234:1544-1549.
22
23          Smith, M; Hopkinson, DA; Harris, H. (1971) Developmental changes and polymorphism
24    in human alcohol dehydrogenase. Ann Hum Genet 34:251-271.
25
26          Soffritti, M; Maltoni, C; Maffei, F; et al. (1989) Formaldehyde: an experimental
27    multipotential carcinogen. Toxicol Ind Health 5:699-730.
28
29          Soffritti, M; Belpoggi, F; Cevolani, D; et al. (2002a) Results of long-term experimental
30    studies on the carcinogenicity of methyl alcohol and ethyl alcohol in rats. Ann NY Acad Sci
31    982:46-69.
32
33          Soffritti, M; Belpoggi, F; Lambertini, L; et al. (2002b) Results of long-term experimental
34    studies on the carcinogenicity of formaldehyde. Ann NY Acad Sci 982:87-105.
35
36          Soffritti, M; Belpoggi, F; Minardi, F; et al. (2002c) Ramazzini Foundation cancer
37    program: history and major projects, life span carcinogenicity bioassay design, chemicals
38    studied, and results. Ann NY Acad Sci 982:26-45.
                                              7-30         DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2          Soffritti, M; Belpoggi, F; Degli Esposti, D; et al. (2005) Aspartame induces lymphomas
 3    and leukemias in rats. Eur J Oncol 10:107-116.
 4
 5          Soffritti, M; Belpoggi, F; Degli Esposti, D; et al. (2006) First experimental demonstration
 6    of the multipotential carcinogenic effects of aspartame administered in the feed to Sprague-
 7    Dawley rats. Environ Health Perspect 114:379-385.
 8
 9          Soffritti, M; Belpoggi, F; Tibaldi, E; et al. (2007) Life span exposure to low doses of
10    aspartame beginnig during prenatal life increases cancer effects in rats. Environ Health Perspect
11    115:1293-1297.
12
13          Speit, G. (2006) The implausibility of systemic genotoxic effects measured by the comet
14    assay in rats exposed to formaldehyde. J Proteosome Res 5:2523-2524.
15
16          Spiteri NJ.  (1982) Circadian patterning of feeding, drinking and activity during diurnal
17    food access in rats. Physiol Behav. Jan;28(l): 139-47. PubMed PMID:7200613.
18
19          Staab, Ca; Lander, J; Brandt, M; et al. (2008) Reduction of S-nitrosoglutathione by
20    alcohol dehydrogenase 3 is facilitated by substrate alcohols via direct cofactor recycling and
21    leads to GSH-controlled formation of glutathione transferase inhibitors. Biochem. J. 413:
22    493-504
23
24          Staats, DA; Fisher, JW; Conolly, RB. (1991) GI absorption of xenobiotics in
25    physiologically based pharmacokinetic models: a two-compartment description. Drug Metab
26    Disp  19(1): 144-8.
27
28          Stanton, ME; Crofton, KM;  Gray, LE; et al. (1995) Assessment of offspring development
29    and behavior following gestational exposure to inhaled methanol in the rat. Fundam Appl
30    Toxicol 28(1): 100-110.
31
32          Starr, TB; Festa, JL. (2003) A proposed inhalation reference concentration for methanol.
33    Regul Toxicol Pharmacol 38:224-231.
34
35          Stegink, LD; Brummel, MC; McMartin, KE; et al.  (1981) Blood methanol concentrations
36    in normal adult subjects administered abuse doses of aspartame. J Toxicol Environ Health 7:
37    281-290.

                                               7-31         DRAFT - DO NOT CITE OR QUOTE

-------
 1          Stegink, LD; Brummel, MC; Filer, LJ; et al. (1983) Blood methanol concentrations in
 2    one-year-old infants administered graded doses of aspartame. J Nutr 113:1600-1606.
 3
 4          Stegink, LD; Filer, LJ Jr; Bell, EF; et al. (1989) Effect of repeated ingestion of aspartame-
 5    sweetened beverage on plasma amino acid, blood methanol, and blood formate concentrations in
 6    normal adults. Metabolism 38(4):357-63.
 7
 8          Steiner, EC; Rey, TD; McCroskey, PS. (1990) Reference guide for Sumusolv. Dow
 9    Chemical Co., Midland, MI.
10
11          Stern, S; Reuhl, K; Soderholm, S; et al. (1996) Perinatal methanol exposure in the rat I:
12    Blood methanol concentration and neural cell adhesion molecules. Fundam Appl Toxicol 34:
13    36-46.
14
15          Sultatos, LG; Pastino, GM; Rosenfeld, CA; et al. (2004) Incorporation of the genetic
16    control of alcohol dehydrogenase into a physiologically based pharmacokinetic model for
17    ethanol in humans. Toxicol Sci. 2004 Mar; 78(1):20-31.
18
19          Svensson, S; Some, M; Lundsj0, A; et al. (1999) Activities of human alcohol
20    dehydrogenase in the metabolic pathways of ethanol and serotonin. Eur J  Biochem 262:324-329.
21
22          Tanner, CM. (1992) Occupational and environmental causes of Parkinsonism. Occup
23    Med 7:503-513.
24
25          Teeguarden, JG; Deisinger, PJ; Poet, TS; et al.  (2005) Derivation of a human equivalent
26    concentration for n-butanol using a physiologically based pharmacokinetic model for n-butyl
27    acetate and metabolites n-butanol and n-butyric acid. Toxicol Sci 85:429-446.
28
29          Teng, S; Beard, K; Pourahmad, J; et al. (2001)  The formaldehyde metabolic
30    detoxification enzyme systems and molecular cytotoxic mechanism in isolated rat hepatocytes.
31    Chem Biol Interact 130-132:285-296.
32
33          Tephly, TR; McMartin, KE. (1984) Methanol metabolism and toxicity. Food Sci Technol
34    12:111-140.
35
36          Tiboni, GM; Giampietro, F; Di Giulio, C. (2003) The nitro oxide synthesis N-omega-
37    nitro-L-arginine methyl ester (L-NAME) causes limb defects in mouse fetuses: protective effect
38    of acute hyperoxia. Pediatr Res 54:69-76.
                                               7-32        DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2          Til, HP; Woutersen, RA; Feron, VJ; et al. (1989) Two-year drinking water study of
 3    formaldehyde in rats. Food Chem Toxicol 27:77-87'.
 4
 5          Thomas, DG; Breslow, N; Gart, JJ. (1997) Trend and Homogeneity Analyses of
 6    Proportions and Life Table Data. Version 2.1. Computers and Biomedical Research 10:373-381.
 7
 8          Thrasher, JD; Kilburn, KH. (2001) Embryo toxicity and teratogenicity of formaldehyde.
 9    Arch Environ Health. 2001 Jul-Aug;56(4):300-ll.
10
11          Tobe, M; Naito, K; Kurokawa, Y. (1989) Chronic toxicity study on formaldehyde
12    administered orally to rats. Toxicology 56:79-86.
13
14          Toth, BA; Wallcave, L; Patil, K;  Schmeltz, I; and Hoffmann, D. (1977) Induction of
15    tumors in mice with the herbicide succinic acid 2,2-dimethylhydrazide.  Cancer Research
16    37:3497-3500.
17
18          Tran, MN; Wu, AH; Hill, DW. (2007) Alcohol dehydrogenase and CAT content in
19    perinatal infant and adult livers: potential influence on neonatal alcohol metabolism. Toxicol
20    Lett. 30;169(3):245-52.
21
22          Trump, BF; McDowell, EM; Harris, CC. (1984) Chemical carcinogenesis in the
23    tracheobronchial epithelium. Environ Health Perspect 55:77-84.
24
25          Tsakiris, T; Giannoulia-Karantana, A; Simintzi, I; et al. (2006) The effect of aspartame
26    metabolites on human erythrocyte membrane acetylcholinesterase activity. Pharmacol Res
27    53:1-5.
28
29          Tsutsui, M; Shimokawa, H; Morishita, T; et al. (2006) Development of genetically
30    engineered mice lacking all three nitric oxide synthases. J Pharmacol Sci 102:147-154.
31
32          Turner, C; Spanel, P; and Smith, D.  (2006) A longitudinal study of methanol in the
33    exhaled breath of 30 healthy volunteers using selected ion  flow tube mass spectrometry, SIFT-
34    MS Physiol Maes. 27:637-648.
35
36          U.S. EPA (Environmental Protection Agency). (1986a) Guidelines for the health risk
37    assessment of chemical mixtures. Federal Register 51(185):34014-34025. Available from:
38    http://www.epa.gov/iris/backgr-d.htm.
                                              7-33         DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2          U.S. EPA (Environmental Protection Agency). (1986b) Guidelines for mutagenicity risk
 3    assessment. Federal Register 51(185):34006-34012. Available from:
 4    http://www.epa.gov/iris/backgr-d.htm.
 5
 6          U.S. EPA (Environmental Protection Agency). (1986c) Rat oral subchronic toxicity study
 7    with methanol. Conducted by Toxicity Research Laboratories, Ltd. TRL No. 032-005.
 8
 9          U.S. EPA (Environmental Protection Agency). (1988) Recommendations for and
10    documentation of biological values for use in risk assessment. Prepared by the Environmental
11    Criteria and Assessment Office, Office of Health and Environmental Assessment, Cincinnati, OH
12    for the Office of Solid Waste and Emergency Response, Washington, DC; EPA 600/6-87/008.
13    Available from: http://www.epa.gov/iris/backgr-d.htm.
14
15          U.S. EPA (Environmental Protection Agency). (1991) Guidelines for developmental
16    toxicity risk assessment. Federal Register 56(234):63798-63826. Available from:
17    http://www.epa.gov/iris/backgr-d.htm.
18
19          U.S. EPA (Environmental Protection Agency). (1994a) Interim policy for particle size
20    and limit concentration issues in inhalation toxicity studies. Federal Register 59(206):53799.
21    Available from: http://www.epa.gov/iris/backgr-d.htm.
22
23          U.S. EPA (Environmental Protection Agency). (1994b) Methods for derivation of
24    inhalation reference concentrations and application of inhalation dosimetry. Office of Research
25    and Development, Washington, DC; EPA/600/8-90/066F. Available from:
26    http://www.epa.gov/iris/backgr-d.htm.
27
28          U.S. EPA (Environmental Protection Agency). (1995) Use of the benchmark dose
29    approach in health risk assessment. Risk Assessment Forum, Washington, DC; EPA/630/R-
30    94/007. Available from: http://cfpub.epa.gov/ncea/raf/recordisplay.cfm?deid=42601.
31
32          U.S. EPA (Environmental Protection Agency). (1996) Guidelines for reproductive
33    toxicity risk assessment. Federal Register 61(212):56274-56322. Available from:
34    http://www.epa.gov/iris/backgr-d.htm.
35
36          U.S. EPA (Environmental Protection Agency). (1997) Guidance on cumulative risk
37    assessment, Part 1 Planning and Scoping. Science Policy Council, Washington, D.C.
38
                                               7-34        DRAFT - DO NOT CITE OR QUOTE

-------
 1          U.S. EPA (Environmental Protection Agency). (1998) Guidelines for neurotoxicity risk
 2    assessment. Federal Register 63(93):26926-26954. Available from:
 3    http://www.epa.gov/iris/backgr-d.htm.
 4
 5          U.S. EPA (Environmental Protection Agency). (2000a) Science policy council handbook:
 6    risk characterization. Office of Science Policy, Office of Research and Development,
 7    Washington, DC; EPA 100-B-00-002. Available from: http://www.epa.gov/iris/backgr-d.htm.
 8
 9          U.S. EPA (Environmental Protection Agency). (2000b) Benchmark dose technical
10    guidance document [external review draft]. Risk Assessment Forum, Washington, DC;
11    EPA/630/R-OO/OO1. Available from: http://www.epa.gov/iris/backgr-d.htm.
12
13          U.S. EPA (Environmental Protection Agency). (2000c) Supplementary guidance for
14    conducting for health risk assessment of chemical mixtures. Risk Assessment Forum,
15    Washington, DC; EPA/63O/R-00/002. Available from: http://www.epa.gov/iris/backgr-d.htm.
16
17          U.S. EPA (Environmental Protection Agency) (2000d) Toxicological review of vinyl
18    chloride. EPA.635R-00/004.
19
20          U.S. EPA (Environmental Protection Agency). (2002a) A review of the reference dose
21    and reference concentration processes. Risk Assessment Forum, Washington, DC; EPA/630/P-
22    02/0002F. Available from: http://www.epa.gov/iris/backgr-d.htm.
23
24          U.S. EPA (Environmental Protection Agency). (2002b) Health assessment of 1,3-
25    butadiene. EPA/600/P-98/001F.
26
27          U.S. EPA (Enviornmental Protection Agency) (2004). Integrated risk information system:
28    vinyl chloride. Available from www.epa.gov/iris/subst/1001 .htm.
29
30          U.S. EPA (Environmental Protection Agency). (2005a) Guidelines for carcinogen risk
31    assessment. Risk Assessment Forum, Washington, DC; EPA/630/P-03/001B. Available from:
32    http://www.epa.gov/iris/backgr-d.htm.
33
34          U.S. EPA (Environmental Protection Agency). (2005b) Supplemental guidance for
35    assessing susceptibility from early-life exposure to carcinogens. Risk Assessment Forum,
36    Washington, DC; EPA/630/R-03/003F. Available from: http://www.epa.gov/iris/backgr-d.htm.
37

                                               7-3 5         DRAFT - DO NOT CITE OR QUOTE

-------
 1          U.S. EPA (Environmental Protection Agency). (2006a) Science policy council handbook:
 2   peer review. Third edition. Office of Science Policy, Office of Research and Development,
 3   Washington, DC; EPA/1 OO/B-06/002. Available from: http://www.epa.gov/iris/backgr-d.htm.
 4
 5          U. S. EPA (Environmental Protection Agency). (2006b) A Framework for Assessing
 6   Health Risk of Environmental Exposures to Children. National Center for Environmental
 7   Assessment, Washington, DC, EPA/600/R-05/093E Available from:
 8   http://cfpub.epa. gov/ncea/cfm/recordisplay.cfm?deid= 158363.
 9
10          U.S. EPA (Environmental Protection Agency). (2006c) Toxic Release Inventory (TRI)
11   On-site and Off-site Reported Disposed of or Otherwise Released (in pounds), for facilities in All
12   Industries, for All Chemicals, U.S., 2003. Available from: http://www.epa.gov/tri. Washington,
13   D.C.
14
15          Vara-Castrodeza, A; Perez-Castrillon, JL; Duefias-Laita, A. (2007) Magnetic resonance
16   imaging in methanol poisoning. Clin Toxicol 45:429-430.
17
18          Ward, KW; Pollack, GM. (1996) Comparative toxicokinetics of methanol in pregnant and
19   NP rodents. Drug Metab Dispos 24:1062-1070.
20
21          Ward, KW; Perkins, RA;  Kawagoe, et al. (1995) Comparative toxicokinetics of methanol
22   in the female mouse and rat. Fundam Appl Toxicol 26:258-264.
23
24          Ward, KW; Blumenthal, GM; Welsch, F; et al. (1997) Development of a physiologically
25   based pharmacokinetic model to describe the disposition of methanol in pregnant rats and mice.
26   Toxicol Appl Pharmacol 145:311-322.
27
28          Weiss, B; Stern, S; Soderholm, SC; et al. (1996) Developmental neurotoxicity of
29   methanol exposure by inhalation in rats. HEI Research Report Number 73.
30
31          Wentzel, P; Rydberg, U; Eriksson, UJ. (2006) Antioxidative treatment diminishes
32   ethanol-induced congenital malformations in the rat. Alcohol Clin Exp Res. 30(10): 1752-60.
33
34          Wentzel, P; Eriksson, UJ. (2006) Ethanol-induced fetal dysmorphogenesis in the mouse
35   is diminished by high antioxidative capacity of the mother. Toxicol Sci 2(2):416-22.
36
37          White, LR; Marthinsen, ABL; Richards, RJ; et al.  (1983) Biochemical and cytological
3 8   studies of rat lung after inhalation of methanol vapour. Toxicol Lett 17:1-5.
                                              7-36         DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2          Wilson, SL; Cradock, MM. (2004) Review: Accounting for prematurity in developmental
 3   assessment and the use of age-adjusted scores. J Pediatr Pschol 29:641-649.
 4
 5          Wu, AH; Kelly, T; McKay, C; et al. (1995) Definitive identification of an exceptionally
 6   high methanol concentration in an intoxication of a surviving infant: methanol metabolism by
 7   first-order elimination kinetics. J Forensic Sci 40(2):315-320.
 8
 9          Yuan J. (1993) Modeling blood/plasma concentrations in dosed feed and dosed drinking
10   water toxicology studies. Toxicol Appl Pharmacol. Mar; 119(1):131-41. PubMed PMID:
11   8470117.
12
13          Zorzano, A; Herrera, E. (1989) Disposition of ethanol and acetaldehyde in late pregnant
14   rats and their fetuses.  Pediatr Res 25(1): 102-6.
                                               7-37        DRAFT - DO NOT CITE OR QUOTE

-------
APPENDIX A. SUMMARY OF EXTERNAL PEER REVIEW AND PUBLIC
                 COMMENTS AND DISPOSITION
                    [Page intentionally left blank]
                             A-1       DRAFT - DO NOT CITE OR QUOTE

-------
           APPENDIX B. DEVELOPMENT, CALIBRATION AND APPLICATION OF A
                                  METHANOL PBPK MODEL
 1          This appendix is adapted from a report prepared for the U. S. EPA under contract by the
 2    Center for Biological Monitoring and Modeling, Battelle Northwest Laboratories (WA 09,
 3    Battelle Project No. 48746, January 31, 2007).

      B.I. SUMMARY
 4          This appendix describes the development, calibration, and approach for application of
 5    mouse, rat, and human PBPK models to extrapolate mouse and rat methanol inhalation-route
 6    internal dose metrics to human inhalation exposure concentrations that result in the same internal
 7    dose (HEC). The human oral methanol dose(s) yielding internal dose(s) equivalent to the mouse
 8    or rat internal dose at the (HED) is also presented.
 9          A PBPK model was developed to describe the blood kinetics of methanol (MeOH) in
10    mice and humans.  The model includes compartments for lung/blood MeOH exchange, liver, fat,
11    and the rest of the body. To describe blood MeOH kinetics, the model employs two saturable
12    descriptions of MeOH metabolic clearance in mice and rats, and one first-order metabolic
13    clearance, and a first-order description of renal clearance (from blood) in humans. Renal
14    clearance is a minor pathway and does not appreciably affect MeOH blood kinetics.
15          This model is a revision of the model reported by Ward et al. (1997), reflecting
16    significant simplifications (removal of compartments for placentae, embryo/fetus, and
17    extraembrionic fluid) and two elaborations (addition of an intestine lumen compartment to the
18    existing stomach lumen compartment and addition of a bladder compartment which impacts
19    simulations for human urinary excretion.), while maintaining the ability to describe MeOH blood
20    kinetics.  The model reported here uses  a single consistent set of parameters; the Ward et al.
21    model employed a number of data-set specific parameters.  Other biokinetic MeOH models that
22    were considered as starting points for the current model also used varied parameters by dataset to
23    achieve model fits to the data. For example, the model of Bouchard et al. (2001) used different
24    respiratory rates and fractional inhalation absorbed for different human exposures.
25          The mouse model was calibrated against inhalation-route blood MeOH kinetic data and
26    verified using intravenous-route blood MeOH kinetic data. The rat model was calibrated against
27    low-dose intravenous data and validated with inhalation-route data. The human model was
28    calibrated against inhalation-route MeOH kinetic data. The models accurately described the
29    inhalation route pharmacokinetics of MeOH. Mouse model simulations of oral- and i.v.-route
30    kinetics compare well to some but not all the experimental data.
31          The MeOH HECs predicted by the model (based on 1,000 ppm inhalation exposure in
32    mice) were >1,000 ppm using either blood AUC or Cmax as the dose metrics. The MeOH HED

                                               B-1         DRAFT - DO NOT CITE OR QUOTE

-------
 1    derived by cross-route extrapolation of this inhalation-route HEC was 110 mg/kg-day, based on
 2    MeOH blood AUC following zero order uptake of MeOH (a constant rate of delivery). Because
 3    of the lack of human data from high-dose exposures, it was not possible to calibrate the model
 4    for inhalation exposures above 1,000 ppm or oral exposures above 110 mg/kg-day.  However,
 5    because the BMD approach was used to estimate an internal experimental animal dose and UFs
 6    were applied to the internal dose, the human model  can be used to back-calculate that internal
 7    dose to an RfC and an RfD below 1,000 ppm or 110 mg/kg-day, respectively.

      B.2.  MODEL DEVELOPMENT
      B.2.1. Model Structure
 8          This model is a revision of the model reported by Ward et al. (1997), reflecting
 9    significant simplifications and two elaborations, while maintaining the ability to describe MeOH
10    blood kinetics in mice, rats, and humans (Figure B-l).  The kidney, pregnancy and the fetal
11    compartment have been removed. The kidney was lumped with the body compartment because
12    the blood:tissue partition coefficients for these tissues were similar. The elaborate time-
13    dependent descriptions of pregnancy were removed because analysis of the available
14    pharmacokinetic data indicates that blood MeOH kinetics in NP and pregnant mice are not
15    different enough to warrant separate descriptions. Because the maternal blood:fetal blood
16    partition coefficients were near  1, there was no need to explicitly model fetal kinetics; they will
17    be equivalent to maternal blood kinetics.  Further supporting data exist for ethanol, which is
18    quite similar to MeOH in its partitioning and transport properties. In rats (Zorzano and Herrara,
19    1989; Guerri and Sanchis 1985), sheep (Brien et al., 1985; Cumming et al., 1984), and guinea
20    pigs (Clarke et al., 1986), fetal and maternal blood concentrations of ethanol are virtually
21    superimposable; maternal to fetal blood ratios are very close to 1, including during late gestation.
22    Also, fetal brain concentrations in guinea pigs (Clarke et al., 1986) were also very similar to the
23    mothers'.
24          In addition to the absolute maternal-fetal concentration similarity noted above, it is
25    common practice to use blood concentrations as an  appropriate metric for risk extrapolation via
26    PBPK modeling for effects in various tissues, based on the reasonable expectation that any
27    tissue:blood differences will be similar in both the test species and humans.  For example, even if
28    the brain:blood ratio was around 1.2 in the mouse or rat, the similar biochemical make-up of
29    brain tissue and blood in rats and humans leads to the expectation that the brain:blood levels in
30    humans (which depend on the biochemical make-up) will also be close to 1.2,  and so the relative
31    "error" that might occur by using blood instead of brain concentration in evaluating the dose-
32    response in rats will be cancelled out by using blood instead of brain concentration in the human.
33    The fact that measured fetal blood levels are virtually identical to maternal levels for methanol
34    (and ethanol) tells us that the rate of metabolism in the fetus is not sufficient to significantly

                                                B-2         DRAFT - DO NOT CITE OR QUOTE

-------
 1   reduce the fetal concentration versus maternal, and use of a PBPK model to predict maternal
 2   levels will give a better estimate of fetal exposure than use of the applied dose or exposure,
 3   because there are animal-human differences in adult PK of MeOH for which the model accounts,
 4   based on PK data from humans as well as rodents.
Inhalation
Exposure
V.

Bladder
KB [1 (human
Urine |>
KLLor
Vma
Km.
I/
JlataKrklita

only)
X-
Vmax
Km2
c
Venous Blood
f

-^
±=^-
2
Kfec^J
Alveolar Air
Lung Blood

Body

Fat

Liver
^Exha



Arterial Blood

tied Air
TKAI \KASorVmAS.KMAS
ntestine^Xr^
-------
 1    fit to data from inhalation and IV exposure, a small rate of elimination from the intestine (lumen)
 2    compartment to feces. (The mouse data could be adequately fit with this rate set to zero,
 3    corresponding to 100% absorption; humans were assumed to have zero fecal elimination, like the
 4    mouse.)  The final models thus include compartments for fat, liver, the rest of the body, bladder
 5    (only used for humans), and lung. The mouse and rat models describe inhalation, oral, and
 6    intravenous route dosing and the human model describes inhalation and oral route dosing and the
 7    rat model includes a non-zero rate  of fecal elimination. Although there is an endogenous
 8    background level of both MeOH and formate (See Section 3.3), the model does not explicitly
 9    describe or account for background levels of MeOH or formate. In this analysis, when non-zero
10    background levels have been measured (in blood), that background was simply subtracted from
11    the concentrations measured during exposure. However, a zero-order rate of infusion could be
12    added to the liver, blood, or stomach compartments to mimic background levels if that was
13    considered necessary.
14          MeOH is well absorbed by the inhalation and oral routes, and is readily metabolized to
15    formaldehyde, which is rapidly converted to formate in both rodents and humans.  Although the
16    enzymes responsible for metabolizing formaldehyde are different in rodents (CAT) and humans
17    (ALD) the metabolite, formate, is the same, and the metabolic rates are similar (Clary, 2003).
18    Most of the published rodent kinetic models for MeOH describe the metabolism of MeOH to
19    formaldehyde as a saturable process but differ in the handling of formate metabolism and
20    excretion (Bouchard et al., 2001; Fisher et al., 2000; Ward et al., 1997; Horton et al., 1992).
21    Ward et al. (1997) used one saturable and one first-order pathway for mice, and Horton et al.
22    (1992) applied two saturable pathways of metabolism to describe MeOH elimination in rats.
23    Bouchard et al. (2001) employed one metabolic pathway and a second pathway described as
24    urinary elimination in rats and humans.  The need for two saturable metabolic pathways in the
25    mouse model was confirmed through simulation and optimization. High exposure (>2,000 ppm
26    MeOH) and low exposure (1,000 ppm MeOH) blood data could not be adequately fit either
27    visually or by more formal optimization without the second saturable metabolic pathway. The
28    optimization approach and results are found below and in the Additional Materials at the end of
29    this appendix.
30          While the PPK model explicitly describes the concentration of methanol, it only
31    describes the rate of metabolism or conversion of MeOH to its metabolites. Distribution and
32    metabolism of formaldehyde is not considered by the model, and this model tracks neither
33    formate nor formaldehyde. (The data that would be needed to parameterize or validate a specific
34    description of either of these metabolites is not available).  Since the metabolic conversion of
35    formaldehyde to formate is rapid (< 1 minute) in all species (Kavet and Nauss, 1990), the MeOH
36    clearance rate may approximate a formate production rate, though this has not been verified.
37    Thus the rate of MeOH metabolism predicted by the model can be used as a dose metric for
38    either or both of these metabolites, but scaling of that metabolic rate metric to humans would
                                               B-4         DRAFT - DO NOT CITE OR QUOTE

-------
 1   require use of an inter-species scaling factor, (BWhUman/BWrodent) '  , to account for the general
 2   expectation of slower clearance of the metabolites in humans.
 3          The model was initially coded in acslXtreme vl .4 and updated in acslXtreme v 2.3
 4   (Aegis Technologies, Huntsville, AL). Most procedures used to generate this report, except those
 5   for the optimization, may be run by executing the corresponding .m files.  The model code
 6   (acslXtreme .csl file) and supporting .m files are available electronically and as text in the
 7   Additional Materials at the end of this appendix. A key identifying .m files associated with
 8   figures and tables in this report is also provided in the Additional Materials.

     B.2.2.  Model Parameters
 9          Physiological parameters such as tissue volumes, blood flows, and ventilation rates were
10   obtained from the open literature (Table B-l). Parameters for blood flow, ventilation, and
11   metabolic capacity were scaled as a function of body weight raised to the 0.75 power, according
12   to the methods of Ramsey and Andersen (1984).

          Table B-l. Parameters used in the mouse and human PBPK models

Body weight (kg)
Mouse
0.03"
Rat
SD | F344
0.2756
Human
70
Source
Measured/estimated
Tissue volume (% body weight)
Liver
Blood arterial
venous
Fat
Lung
Rest of body
5.5
1.23
3.68
7.0
0.73
72.9
3.7
1.85
4.43
7.0
0.50
73.9
2.6
1.98
5.93
21.4
0.8
58.3
Brown etal. 1997
Calculatedc
Flows (L/hr/kg0'75)
Alveolar ventillation''
Cardiac output
25.4
25.4
16.4
16.4
16.5
24.0
Perkins et al. 1995a; Brown et al. 1997;
U.S. EPA, 2004
Percentage of cardiac output
Liver
Fat
Rest of body
25.0
5.0
70.0
Biochemical constants8
VmaxC (mg/hr/kg075)
Km (mg/L)
Vmax2C (mg/hr/kg075)
Km2 (mg/L)
K1C (BW025/hr)
KLLC (BW025/hr/
19
5.2
3.2
660
NA
NA
25.0
7.0
68

5.0 0
6.3 NA
8.4 22.3
65 100
NA
NA
22.7
5.2
72.1
1s*
, saturable
order
NA 33.1
NA 23.7
NA
NA
0.0373 0.0342
95.7 NA
Brown etal. 1997
Calculated

Fitted
                                               B-5
DRAFT - DO NOT CITE OR QUOTE

-------
Oral absorption
VmASC (mg/hr/kg075)
KMASC (mg/kg)
KSI (hr1)
KAI (hr"1)
Kfec (hr1)
1830
620
2.2
0.33
0
5570
620
7.4
0.051
0.029
377
620
3.17
3.28
0
Mouse and rat fitted (mouse and human
KMASC assumed = rat); other human
values are those for ethanol from Sultatos et
al. (2004), with VmASC set so that for a
70-kg person VmAS/KM = the first-order
constant of Sultatos et al.
Partition coefficients
Liver:Blood
Fat:Blood
Blood:Air
Body:Blood
Lung:Blood
Bladder time-constant
(KBL, hr-1/
Inhalation fractional
availability (%)
1.06
0.083
1350'
0.66
1
1.06
0.083
1350
0.66
1
NA
0.665
0.20
0.583*
0.142
1626
0.805
1.07
0.564 0.612
0.866*
Ward et al., 1997; Fiserova-Bergerova and
Diaz, 1986
Horton et al., 1992; Fiserova-Bergerova and
Diaz, 1986
Rodent: estimated; human: Fiserova-
Bergerova and Diaz, 1986 (human "body"
assumed = muscle)
Fitted (human)
Rodent: fitted; human
Ernstgard et al., 2005
1
2
3
4
5
6
7
NA - Not applicable for that species
"Both sources of mouse data report body weights of approximately 30 g
*The midpoints of rat weights reported for each study was used and ranged from 0.22 to 0.33 kg
The volume of the other tissues was subtracted from 91% (whole body minus a bone volume of approximately 9%) to get
the volume of the remaining tissues
''Minute ventilation was measured and reported for much of the data from Perkins et al. (1996) and the average alveolar
ventilation (estimated as 2/3 minute ventilation) for each exposure concentration was used in the model. When ventilation
rates were not available, a mouse QPC (Alveolar Ventilation/BW°75) of 25.4 was used (average from Perkins et al., 1995a).
The QPC used to fit the human data was obtained from U.S. EPA (2004). This QPC was somewhat higher than calculated
from Brown etal.  (1997) (-13 L/hr/kg075)
eVmax, Km, and Vmax2, Km2 represent the two saturable metabolic clearance processes assumed to occur solely in the liver.
The Vmax used in the model = VmaxC (mg/kg°75-hr) XBW075. K1C is the first-order loss from the blood for human
simulations that represents urinary elimination. Allometric scaling for first-order clearance processes was done as previously
described (Teeguarden et al., 2005); The Kl used in the model= K1C / BW°25
fKLLC - alternate human first-order metabolism rate (used only when VmaxC =  Vmax2C = 0)
sHuman oral simulations used a zero order dose rate equal to the mg/kg.day dose
*Human liverblood estimated from correlation to (measured) fatblood, based on data from 28 other solvents
7Rat partition coefficient used for mice as done by Ward et al. (1997)
JKBL - a first-order rate constant for clearance from the bladder compartment, used to account for the difference between
blood kinetics and urinary excretion data as observed in humans
*For human exposures, the fractional availability was from Sedivec  et  al. (1981), corrected for the fact that alveolar
ventilation is 2/3 of total respiration rate

        Mouse model partition coefficients were used as  reported (liver, fat, blood:air) or

estimated (lung, body).  The mouse body compartment partition coefficient was set

approximately equal to the measured value for muscle (Ward et al., 1997). The mouse lung

partition coefficient was assumed to be 1.0, similar to the liver partition coefficient.  This

parameter has no numerically significant impact on modeled blood dose metrics.

        Human partition coefficients were reported by Horton et al. (1992), but were in fact

measured in rat tissues. The reported rat fat partition coefficient was considerably closer to unity

than reported for MeOH or ethanol by other researchers  (Ward et al.,  1997; Pastino  and Conolly,
                                                   B-6
                                                            DRAFT - DO NOT CITE OR QUOTE

-------
 1   2000) and assumed to be in error. Human partition coefficients were obtained from Fiserova-
 2   Bergerova and Diaz (1986).

     B.2.3. Mouse Model Calibration

       B. 2.3.1.  Inhalation-Route Calibration

 3          For purposes of conducting interspecies extrapolations of MeOH dosimetry, the
 4   inhalation route was the most important route requiring calibration for the mouse model. The
 5   critical endpoint and NOEL, which are the basis for the HEC estimation, are from inhalation-
 6   route studies.  The ability to predict blood MeOH concentrations from inhalation exposures was
 7   therefore  a priority. Pharmacokinetic data from other routes, i.v. and oral, were used to verify
 8   clearance terms derived by fitting to the inhalation data or to estimate a MeOH oral uptake rate
 9   constants. Holding other parameters constant, the mouse PBPK model was calibrated against
10   inhalation-route blood pharmacokinetic data (Figure B-2) by fitting five parameters: Michaelis-
11   Menten constants for one high affinity/low capacity and one low-affinity high-capacity enzyme
12   and the inhalation  fractional availability term.
                   10,000
                    1,000
                 I
                 o
                 Oj
                     100
                      10
                                                       — 1000 pprn sim
                                                       — 2000 pprn sim
                                                       — 5000 ppm sim
                                                       — 7500 ppm sim
                                                       — 10000 ppm sim
                                                         15000 pprn sim
                                                       A 1000 pprn data (GD6 & 10)
                                                       • 2000 ppm data (GD6 &. 10)
                                                       • 5000 ppm data (GD6 & 10)
                                                       + 7500 ppm data (GD6 & 10)
                                                       A 10000 pprn data (GD6 & 10)
                                                       V 10000 ppm data (GD7)
                                                       n 15000 ppm data (GD6 & 10)
                                                12    15
                                                  Time (hr)
                                                             18
                                                                   21
                                                                         24
                                                                               27
13
14
15
       Figure B-2. Model fits to data sets from GD6, GD7, and GD10 mice for
       7-hour inhalation exposures to 1,000-15,000 ppm MeOH. Maximum
       concentrations are from Table 2 in Rogers et al. (1993). The complete data set for
       GD7 mice exposed to 10,000 ppm is from Rogers et al. (1997) and personal
       communication (Additional Materials).  Symbols are concentration means of a
       minimum of n = 4 mice/concentration. Default ventilation rates (Table B-l) were
       used to simulate these data.

       For these mouse simulations, pulmonary ventilation was set to 25.4 (L/hr/kg0'75), the
average value measured by Perkins et al. (1995a), which is similar to the value of 29 (L/hr/kg0'7
reported in Brown et al. (1997). Where ventilation rates were reported for individual exposure
                                                B-7
                                                       DRAFT - DO NOT CITE OR QUOTE

-------
 1    concentrations by Perkins et al. (1995a), they were used directly in the model and a notation was
 2    made in the figure legend. Reported ventilation rates varied from 592 to 857 L/kg x 8 hr,
 3    depending on exposure concentration (Perkins et al., 1995a). Adjusting these values to 2/3 total
 4    (for alveolar ventilation) and allometrically scaling by BW°'75, values used in the model for these
 5    exposures ranged from 20.5 to 29.7 (L/hr/kg0'75) (See Table B-l). A fractional availability of
 6    73% of alveolar ventilation was visually optimized to best describe the inhalation-route blood
 7    MeOH pharmacokinetic data. This percentage of uptake for inhalation exposures is similar to
 8    values reported for other alcohols in rodents (Teeguarden et al., 2005), but considerably lower
 9    than the value reported by Perkins et al. (1995a) of 126% of alveolar ventilation (85% of total
10    ventilation).
11          The calibrated model predicted blood MeOH concentration time-course agreed well with
12    measured values in adult mice in the inhalation studies of Rogers et al. (1997, 1993) (Figure B-
13    2), and Perkins et al.  (1995a), as well as in NP and early gestation (GD8) mice of Dorman et al.
14    (1995) (Figure B-3).  Parameter values used in the calibrated model are given in Table B-l.
                8,000-
                7,000-
              J 6,000-
              £ 5,000-
              O 4,000-
              4)
                3,000-
              •o
              0
              5 2'°°°-
                1,000-
                    0-
+ 2,500 ppm NP mouse
A 5,000 ppm NP mouse
o 10,000 ppm  NP mouse
D 6 hr, 15,000 ppm GD 8 mouse (Dorman) J
                       0
 15      20
  Time (hr)
30
35
          Figure B-3. Simulation of inhalation exposures to MeOH in NP mice from Perkins
15        et al. (1995a) (8-hour exposures) and Dorman et al. (1995), (6-hour exposures). Data
16        points represent measured blood MeOH concentrations and lines represent PBPK model
17        simulations. Note: data was obtained using Digitizlt (Sharlt! Inc. Greensburg, PA) to
18        digitize data from Figure 2 of Perkins et al.  (1995a) and Figure B-2 from Dorman et al.,
19        (1995). Default ventilation rates (Table B-l) were used to simulate the Dorman data.  The
20        alveolar ventilation rate for each data set from Perkins et al.  (1995) was set equal to the
21        measured value reported in that manuscript. For the 2,500, 5,000, and 10,000 ppm exposure
22        groups, the alveolar ventilation rates were 29, 24, and 21 (L/hr/kgO.75), respectively. The
23        cardiac output for these simulations was set equal to the alveolar ventilation rate.
                                                            DRAFT - DO NOT CITE OR QUOTE

-------
       B. 2.3.1.  (should be B. 2.3.2) Oral-Route Calibration
 1          The mouse model was calibrated for the oral route by fitting the rate constants for oral
 2    uptake of MeOH. Calibration of the oral route was not required for interpretation of the critical
 3    toxicology studies.  This exercise was undertaken to estimate the rate constants for oral uptake so
 4    it could be used to make dose-route extrapolations for calculating human oral-route exposures
 5    equivalent to mouse exposures at the NOEL.
 6           Ward et al. (1997) described MeOH uptake as the sum of a fast and slow process (two
 7    rate constants), with a fraction of the administered dose attributed to each process. The rate
 8    constants and the fraction of the dose attributed to each process were varied to describe oral-
 9    route blood MeOH kinetics for each GD. For instance, the fraction of the total oral dose
10    assigned to the fast absorption process varied from 54 to 71%, depending on the data set. An
11    alternative approach with uptake attributed to stomach and intestine, which allows for greater
12    flexibility in fitting the data (Staats et al., 1991), was compared to a simpler one utilizing a single
13    rate of uptake.  In both the current model and the model of Ward et al. (1997), orally ingested
14    MeOH was assumed to be 100 % absorbed.
15          Initially, a single oral absorption rate constant (KAS, hr"1) was fitted to oral-route blood
16    MeOH kinetics reported by Ward et al. (1997, 1995). Using these data, an average KAS
17    (0.62 hr"1) was estimated that provides adequate fits to MeOH blood kinetics following 2,500
18    mg/kg dose in NP and GDIS mice and 1,500 mg/kg in GD8 mice up to ~8 hours. At later time
19    points, however, a model using a single oral uptake rate constant consistently under predicts
20    blood concentrations of MeOH (results not shown).  Fits were improved by using the two
21    compartment GI tract model (Figure B-4). However, when fitting the  oral data in rats, it was
22    found that the fits were significantly improved if the uptake from the stomach was treated as a
23    saturable process. Vmax (VMASC) was  scaled as BW°'75, as is done for other Vmaxs, and the Km
24    (KMASC) was scaled as BW1 to reflect that the variable used is the total amount in the stomach,
25    whose volume is  expected to scale with BW1. For the mouse, model fits were not significantly
26    improved when KMASC was allowed to vary (change from the value fitted to rats,  1830 mg/kg),
27    so it was kept at the rat value.
28          Using the two-compartment oral absorption model and adjusting only the absorption
29    parameters resulted in a good fit to the lower oral dose (1,500 mg/kg) (Dorman et al., 1995), but
30    consistently under-prediction of the 2,500 mg/kg oral dosing blood levels (Ward et al., 1997).
31    When the metabolic constants (VmaxC values) were decreased, the data from the higher dose
32    were fit, but the fit of the data for the 1,500 mg/kg dose was lost (see Additional Materials,
33    Figure B-19). Also, when using the lower clearance required to fit the data of Ward et al. (1997),
34    the inhalation data of Rogers et al. (1993) could no longer be fit by the model (see Additional
35    Materials, Figure B-20).  The two-compartment GI tract approach (with parameters that better  fit
36    the low dose  data) was retained in the model and used for all final mouse oral route simulations.
37
                                               B-9          DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
13
14
15
16
17
18
19
        o
       O
       m
     3,500-

     3,000-

     2,500-

     2,000-

     1,500-
        o 1,000-
        OD
            500-
              0-1
• Oral 1500 mg/kg GD 8 mouse
• Oral 2500 mg/kg NP mouse
n Oral 2500 mg/kg GD IS mouse
— 1500 mg/kg simulation
— 2500 mg/kg simulation
                                            10             15
                                                Time (hr)
                                                                    20
                             25
    Figure B-4. Oral exposures to MeOH in pregnant mice on GD8 (Dorman et al., 1995)
    or NP and GDIS. Data points represent measured blood concentrations and lines represent
    PBPK model estimations for NP mice.
       Source: Ward et al., (1997).

 B. 2.3.1. (B2.3.3) Intravenous Route Simulation
       The parameterization of MeOH clearance (high-and-low affinity metabolic pathways)
was verified by simulation of data sets describing the intravenous-route pharmacokinetics of
MeOH. MeOH blood kinetics data in NP mice are only available for a single i.v. dose of
2,500 mg/kg (Ward et al., 1997). MeOH blood kinetics are also reported in GDIS mice
following administration of a broader range of doses: 100, 500, and 2,500 mg/kg.  Because
MeOH kinetics appear similar for NP and pregnant mice after administration of 2,500 mg/kg
prior to 20 hours, the model is expected to fit data for both pregnant and NP mice using the same
set of parameters, and hence, data  for both life stages were used to verify metabolic clearance of
MeOH.
       Initial blood concentrations of MeOH following i.v. administration were not proportional
to administered dose in the data from Ward et al. (1997).  Initial concentrations were
approximately 1.5-fold lower in the 100 mg/kg dose group than expected if a dose-independent
volume of distribution (VD) is assumed (Figure B-5A). Initial blood concentrations were,
however, proportional to administered dose between 2,500 and 500 mg/kg.  Two possible
explanations were then considered:
   1) the VD, which is not impacted by any other PBPK parameters and is only determined by
      the biochemical partitioning properties of MeOH, is twofold lower at 100 mg/kg than at
      the higher concentrations, while the VD at 500 and 2,500 mg/kg are exactly as predicted
      by the PBPK model without adjustment; or
                                              B-10
                                                     DRAFT - DO NOT CITE OR QUOTE

-------
 1       2) a dilution error occurred during preparation of the 100 mg/kg dosing solution used by
 2          Ward etal. (1997).
    10,000




  ~  1,000
  _l
  ~~-
  01
  E
              o
              Oj
              •o
              o
              0
                   100
                    10
                   1,000
                     100
                 O
                 a
                 o    10
                              - IV 2500 simulation
                              - IV 500 simulation
                              — IV 100 simulation
                              • IV 2500 NP mouse
                              a IV 2500 GD 8 mouse
                              D IV 2500 GD IS mouse
                              i IV 500 GD 18 mouse
                              • IV 100 GD 18 mouse
                                               10
                                                           15
                                                  Time (hr)
                                                                                   A
                                                                        20
                                                        '-— IV 200 simulation  }
                                                        — IV 100 simulation
                                                          IV 200 GD 18 mouse I
                                                                                    25
                                                                        B

                             0.5
                                          1,5
                                                 2     2,5
                                                 Time (hr)
                                                                   3.5
                                                                                4.5
 3
 4
 5
 6
 7
 8
 9
10
11
Figure B-5. Mouse intravenous route MeOH blood kinetics. A) MeOH was
infused over 1.5 minutes into female CD-I mice at target doses of 100 (circles),
500 (triangles) or 2,500 (squares) mg/kg. Mice were NP, GD9  or GDI8 at the
time of dosing. Data points represent measured blood concentrations and lines
represent PBPK model simulations. B) Comparison of the 100 mg/kg dose data
(points) and PBPK model simulations assuming a 100 mg/kg dose as reported
(solid line) or to a presumed 200 mg/kg dose (dashed lines). Note: the 24-hour
time point data from the 500 mg/kg NP and GD 9 mice are below the reported
detection limit (2 |ig/ml) and so are not shown.

Source:  Adapted from Ward et al. (1997).
 3          To account for this unexpected nonproportionality, Ward et al. (1997) used higher
 4   partition coefficients for placenta and embryonic fluid and lower Vmax for the metabolism of
                                               B-11         DRAFT - DO NOT CITE OR QUOTE

-------
 1   MeOH for the 100 and 500 mg/kg doses than for the 2,500 mg/kg dose. These adjustments to
 2   partition coefficients effectively change the volume of distribution.  However, the PBPK model
 3   obtained with measured partition  coefficients and otherwise calibrated to inhalation data, as
 4   described above, was capable of simulating both the 500 and 2,500 mg/kg data without adjust-
 5   ment or varying parameters between those 2 doses. Further, the data at the nominal dose of 100
 6   mg/kg could also be adequately fit without other parameter adjustment simply by simulating a
 7   dose of 200 mg/kg (dotted line, Figure B-5B). If the relative difference in the VD at 100 mg/kg
 8   was not a round number and/or the apparent value at 500 mg/kg was intermediate between 100
 9   and 2,500 mg/kg, the possibility of a dose-related variation in this or other parameters would be
10   given more weight. Also it seems unlikely that the volumes of distribution would be different in
11   these animals solely based on differing exposure concentrations.  But the far simpler and more
12   likely explanation for the observed pattern of VD seems to be a dosing error (2).  Therefore, a
13   single set of parameters was retained which describe the 2,500 and  500 mg/kg doses rather than
14   adjust parameters to fit a data set  (the  100 mg/kg dose group) that appeared inconsistent and may
15   be the result of a simple experimental  error rather than attributable to a dose dependence.
16       Thus, high- (2,500 mg/kg) and mid-dose (500 mg/kg) intravenous-route pharmacokinetic
17   data were used to validate the parameters calibrated from the inhalation studies for the metabolic
18   clearance of MeOH.  Metabolic constants reasonably predict blood MeOH kinetics following a
19   2,500 mg/kg dose in NP animals until 12 hours postexposure, but under predict blood MeOH in
20   GD9 and GDIS mice at 8 hours of exposure and beyond, and under-predict levels in both NP and
21   pregnant mice at 15 hours and beyond. At this high-dose,  where blood kinetics of MeOH were
22   reported in NP, GD9, and GDIS mice, the data for the GDIS mice was inconsistent with the
23   GD9 and NP animals. The GD9 data  at 12 hours appears inconsistent with the NP data, but then
24   the 2 are nearly identical again at 15 hours, so it is not clear if that difference at 12 hours is real
25   or just due to experimental variability. Blood levels of MeOH were -500 mg/L in GDIS  mice at
26   24 hours, but were nondetectable  after 18 hours  in the other groups (detection limit 2  mg/L).
27   Blood concentrations were accurately  predicted  following administration of 500 mg/kg MeOH
28   (Figure B-5A). The model predictions did not match the 100 mg/kg data unless one assumed an
29   error in dose preparation, as described above (Figure B-5B). The calibration of the MeOH
30   PBPK model is consistent with both the available inhalation and  oral-route data.

       B.2.3.1. (B2.3.4)TotalMethanolMetabolic Clearance
31          Quantifying production of formaldehyde following MeOH exposure for use as an
32   alternative dose metric is  of particular interest because formaldehyde is also undergoing toxicity
33   assessment. However, it is important  to understand that because  the model was developed to
34   describe blood MeOH kinetics, metabolism of MeOH to neither formaldehyde nor formate is
35   specifically described; the model  tracks neither of these metabolites.  While the metabolic
36   clearance of MeOH described by  the model may be presumed to  equate with formaldehyde

                                               B-12        DRAFT - DO NOT CITE OR QUOTE

-------
 1    production, this metabolic flux simply leaves the computational model system without specific
 2    attribtution. Since the metabolic conversion of formaldehyde to formate is rapid in all species
 3    (< 1 minute) (Kavet and Nauss, 1990), the MeOH clearance rate may approximate a formate
 4    production as well as a formaldehyde production rate, though this has not been verified.
 5          Thus, production of formaldehyde or formate following exposure to MeOH can only be
 6    estimated by summing the total amount of MeOH cleared by metabolic processes. If used, this
 7    metric of formaldehyde or formate dose should be adjusted by an inter-species scaling factor
 8    (SF = [BWhuman/BWrodent]0'25) to adjust for expected species differences in the clearance of these
 9    two metabolites (this is a default factor based on the generally accepted assumption that total
10    metabolism scales as BW°75 and hence clearance per BW scales as 1/BW0'25).  The rate of
11    MeOH clearance may roughly be equated to the total amount of metabolites produced. Values of
12    total MeOH clearance as a function of exposure in mice and humans are presented in the
13    Additonal Materials (Tables 7-9).

       B. 2.3.1. (B.2.3.5) Formal Optimization of Mouse Model Parameters
14          Formal  optimization of five parameters (inhalation fractional availability and the Vmax
15    and Kms for high and low affinity MeOH metabolism) was attempted using optimization
16    routines in acslXtreme v2.01.1.2.  Under the best circumstances, formal optimizations offer the
17    benefit of repeatability and confirmation that global optima have not been significantly missed
18    by user-guided visual optimization. Incorporating judgments regarding the value of specific data
19    sets, while possible when visually fitting, is more difficult when using optimization routines.
20    This is an important distinction between these approaches for this modeling exercise.
21          The mouse inhalation route NOEL was less than 1,000 ppm MeOH. The model is
22    calibrated against inhalation-route data because of the importance of this exposure route in the
23    assessment. Unfortunately, the vast majority of the MeOH data came from much higher
24    exposure concentrations. As expected, various attempts at formal optimization lead to improved
25    fits for some but never all data  sets. This is to be expected when there is significant variability in
26    the underlying  data. Various data-weighting schemes were included to improve overall
27    optimization while maintaining a good fit to the lowest concentration (1,000 ppm) data. In the
28    end, formal optimization provided no significant improvement over the fractional availability
29    and metabolic parameter values obtained by visual optimization, so these were retained in the
30    final version of the model.
31          Further details on the approach and  results from the formal optimization are found in the
32    Additional Materials in outline format with supporting figures. More complete documentation
33    was not developed because the products of the optimizations were not used in the final model.
34    The documentation is intended only to demonstrate that appropriate optimizations were
35    conducted and  what the results of those  optimizations were.


                                               B-13         DRAFT - DO NOT CITE OR QUOTE

-------
      B.2.4. Mouse Model Sensitivity Analysis
 1          An evaluation of the importance of selected parameters on mouse model estimates of
 2    blood MeOH AUC was performed by conducting a sensitivity analysis using the subroutines
 3    within acslXtreme.  Files for reproducing the sensitivity analysis are available in the model as
 4    described in the additional materials.  The analysis was conducted by measuring the change in
 5    model output corresponding to a 1% change in a given model parameter when all other
 6    parameters were held fixed. A normalized sensitivity coefficient of 1 indicates that there is a
 7    one-to-one relationship between the fractional change in the parameter and model output; values
 8    close to zero indicate a small effect on model output. A positive value for the normalized
 9    sensitivity coefficient indicates that the output and the corresponding model parameter are
10    directly related while a negative value indicates they are inversely related.
11          Sensitivity analyses were conducted for the inhalation and oral routes. The
12    inhalation-route analysis was conducted under the exposure conditions of Rogers and Mole
13    (1997) and Rogers et al. (1993), 7-hour inhalation exposures at the NOEL concentration of
14    1,000 ppm.  The oral route sensitivity analysis was conducted for an oral dose of 1,000 mg/kg.
15          Parameters with sensitivity coefficients less than 0.1 are not reported. The parameters
16    with the largest sensitivity coefficients for the inhalation route at 1,000 ppm were pulmonary
17    ventilation, VmaxC, and partitioning to the body compartment (Figures B-6 [metabolism]  and B-7
18    [flows and partition coefficients]). MeOH AUC  was also sensitive to KM2 and  VmaxC. The
19    sensitivity coefficient for pulmonary ventilation increases from 1  to -1.75 during the exposure
20    period as metabolism begins to saturate.  The sensitivity coefficient is 1 for concentrations 100
21    ppm MeOH or less or when hepatic clearance is nonlimiting.
22          Oral-route mouse blood MeOH AUC was sensitive to the rate constants for uptake.
23    Blood AUC was most sensitive to the first-order rate constant for uptake from the stomach, KAS,
24    during the first hour after exposure, becoming less important over time (Figure B-8). Blood
25    MeOH AUC was also modestly sensitive to KAI, and KSI, the rate constants for uptake from the
26    intestine and transfer rates between compartments, respectively.
27
                                               B-14         DRAFT - DO NOT CITE OR QUOTE

-------



c
d>
0
o

'fn
0)





Oc
.O
*i t;

2C
.0
o
-o
3C
^A
lfm

r^~ r\m«_ ~x
r /
\
%x
\
V
* V, %
"" * ^
«^_ 	 ./—»*—• VmaxC • - -
                                          6       8      10
                                             Time (hr)
                                                       12
14
16
         Figure B-6. Mouse model inhalation route sensitivity coefficients for metabolic
1        parameters. Sensitivity coefficients calculated for an exposure of 1,000 ppm MeOH are
2        reported for blood MeOH AUG.  Note: Km, Vmax refer to the high-affinity, low-capacity
3        pathway and Km2, Vmax2 refer to the low-affinity, high-capacity pathway.


d>
0
it
0)
o
o

'fn
o>



t
3C
.0
2C

•\ <^


0
U.O
n

n
-0.5
(

S *~r*. 	
*"
QPC"
s
s
/
\

i QLC-J"^ 	

D 2 4 6 8 10 12 14 1
Time (hr)
1
2
Figure B-7. Mouse model inhalation route sensitivity coefficients for flow rates (QCC:
cardiac output; QPC: alveolar ventilation), and partitioning to the body (PR)
compartment are reported for blood MeOH AUC. Sensitivity coefficients calculated for
an exposure to 1,000 ppm MeOH.
                                            B-15
                                                DRAFT - DO NOT CITE OR QUOTE

-------
             0)
            'o
            1
             8
             (A
             0)
                 0.8
                 0.6
                 0.4
                 0.2
                   0
                -0.2
                -0.4
                -0.6
\
 VmASC
                        KAI
                                    KSI
                       KMASC
                                            6       8      10
                                                Time (hr)
                                                                    12
                                                     14
16
1        Figure B-8. Mouse model sensitivity coefficients for oral exposures to MeOH. The
2        sensitivity of blood MeOH AUC to oral absorption rate constants (KAS: stomach; KAI:
3        intestine; KSI: transfer between compartments) is reported.
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
    B.2.5. Mouse Drinking Water Ingestion Pattern
           To simulate exposures of mice via drinking water under bioassay conditions, an ingestion
    pattern first used by Keys et al. (2004), based on data from Yuan (1993) was used. The pattern
    specifies a fraction of percent of total daily ingestion consumed in each half-hour interval. The
    first interval was shifted to correspond to the beginning of the active (dark) period, for
    consistency with patterns used for humans and rats.  A Table function was used in acslXtreme to
    interpolate an instantaneous rate between the measured  (30-min) values, with normalization so
    that the 24-hour integral equals 100%.  The daily pattern is shown in Figure B-9A and the
    resulting blood concentration for a mouse exposed for 6 days per week (2100 mg/kg)  is  shown in
    Figure B-9B.
                                              B-16
                                                            DRAFT - DO NOT CITE OR QUOTE

-------
 ^-, 1,500-,
 _l

 cn

 E 1,200-
  c
  o
 S3
  ID
 b
  c
  •D
  U
  C
  o
  U

 ~0
  o
 _0

 CD
900-
600-
300-
                                                   i— —r
                                                  18   20
                                                      22   24
                                 Time (h)
                                                       B
                 24
                   48
72      96

 Time (h)
120
144
168
Figure B-9. Mouse daily drinking water ingestion pattern (A) and resulting

predicted blood concentration for a 6 d/wk exposure (B). Mouse drinking

water exposures were simulated by multiplying the fractional rate (1/h) as a

function of clock time by the daily total dose ingested (mg) to obtain a rate of

addition of methanol into the stomach lumen compartment (mg/h).


Source: Yuan (1993); Keys et al. (2004)
                                 B-17
                                        DRAFT - DO NOT CITE OR QUOTE

-------
      B.2.6. Rat Model Calibration
 1          The model was initially calibrated-to-fit data from intravenous, inhalation, and oral
 2    exposures in Sprague-Dawley (SD) rats using the 100 and 2500 mg/kg intravenous (IV) data
 3    provided in the command file of Ward et al. (1997). Holding other parameters constant, the rat
 4    PBPK model was calibrated against the Ward et al. (1997) IV-route blood pharmacokinetic data
 5    (Figure B-10) by fitting Michaelis-Menten constants for one high affinity/low capacity and one
 6    low-affinity, high-capacity enzyme, using the optimization routines in acslXtreme v2.3. Also
 7    shown for comparison in Figure B-10A are the 100 mg/kg IV data of Horton et al. (1992),
 8    obtained using Fischer 344  (F344) rats (data extracted from figures using Digitizlt), with a model
 9    simulation (heavy red line)  which differs from that for the SD rat only due to the predicted effect
10    of know body weight differences. While the fit to the Ward et al. data for SD rats is excellent,
11    especially for the lower dose, the rate of clearance is over-predicted for the F344 rat when
12    parameters fit to SD rat data are used. The 100  mg/kg IV data, with an alternate simulation for
13    the F344 rat obtained with distinct parameters (see below) is expanded in Figur B-10B,
14    emphasizing the difference  in clearance between the two strains.
15          We then attempted to fit the model to the inhalation data of Horton et al.  (1992) by
16    adjusting only the inhalation fractional uptake (FRACIN). The results, shown in Figure B-11A,
17    are clearly poor. While the  model does  match the uptake portion of the inhalation data for the
18    1200 and 2000 ppm exposures, it under-predicts the peak concentration reached at 200 ppm.
19    Further, the post-exposure clearance predicted by the model is much more rapid than indicated
20    by the data, as occurred with the IV kinetics (Figure B-10). (Since the peak concentration for the
21    2000 ppm inhalation exposure actually occurred at 7 hr, we also simulated a 7-hr exposure,
22    shown by the thin black line. The result indicates that the data are more consistent with and
23    better predicted by the longer exposure duration, but clearance is still over-predicted post-
24    exposure.) Therefore we concluded that the data show a clear strain difference in metabolism,
25    and should support  at least a partially independent set of parameters for SD and F344  rats.
26          We then combined the 100 mg/kg IV and inhalation data of Horton et al. (for F344 rats)
27    and attempted to simultaneously identify the four metabolic parameters (Vmax and Km for two
28    pathways) and FRACIN.  However when this was done the resulting values for the two Km's
29    were ~ 90 ± 50 mg/L and 70 ± 40 mg/L (Km and Km2, respectively), which are clearly
30    indistinguishable from a statistical standpoint.  If instead the Km's were fixed at the more distinct
31    values identified from the SD rat IV data (6.3 and 65 mg/L), the optimization routine tended to
32    set the Vmax associated with the lower Km to zero. Thus the F344 rat data of Horton et al.
33    (1992) appear to be most consistent with a single metabolic pathway, even though the observed
34    concentrations spanned almost 2 orders of magnitude. Therefore those data (including the 100
35    mg/kg IV data) were simultaneously fit by adjusting a single Vmax and Km, along with the
                                               B-18         DRAFT - DO NOT CITE OR QUOTE

-------
1   inhalation fraction, FRACIN, with the second metabolic pathway set to zero (Figures B-10B and

2   B-11B).

               10,000
            en
            I
            O
            
-------
    100
     10
 I
 o
 v
 •o
 o
 o
    0.1
          — 2000 ppm simulation     — 1200 ppm simulation
          — 200 ppm simulation      o 2000 ppm data
          * 1200 ppm data          A 200 ppm data
          — 2000 ppm, 7-hr simulation
                                6       8      10
                                    Time (hr)
        12
14      16
    100
     10
 I
 O
 0!
 o
 o
    0.1
                                                               B
         l	IN.	
         f— 2000 ppm simulation     — 1200 ppm simulation
          — 200 ppm simulation      D 2000 ppm data
          *• 1200 ppm data          A 200 ppm data
         (j- 2000 ppm, 7-hr simulation	
                                        8
                                    Time (hr)
10
14      16
Figure B-ll. Model fits to data sets from inhalation exposures to 200
(triangles), 1,200 (diamonds), or 2,000 (squares) ppm MeOH in male F-344
rats. (A) Model fits with metabolic parameters set to values obtaine from IV data
for Sprague-Dawley rats, with only the inhaled fraction (FRACIN) adjusted. (B)
Model fits obtained by fitting metabolic parameters (Vmax and Km) for a single
pathway, along with FRACIN, to these data as well as the 100 mg/kg IV data
from F-344 rats (Figure B-9B).

Symbols are concentrations obtained using Digitizlt!.  Lines represent PBPK
model fits. As the 7-hour data point at 2,000 ppm is higher than the 6-hour data
point (more evident on a linear scale) and appears more consistent with a 7-hour
exposure, a model simulation for a 7-hour exposure at 2,000 ppm is also shown
(lighter line).

Source:  Horton et al. (1992)
                                   B-20
DRAFT - DO NOT CITE OR QUOTE

-------
 1          Model simulations of the F344 rat data with the F344-specific parameters are shown in
 2    Figure B-10B (heavy red line) and Figure B-lIB. Unfortunately we were unable to
 3    simultaneously fit the inhalation data for all exposure levels, although a wide range of metabolic
 4    saturation (Km) values were tested.  We could obtain a better fit of the high-concentration data
 5    by constraining FRACIN to a higher value, for example, but then the fits to the lower
 6    concentration data were compromised (not shown). Examining Figure 2 of Horton et al. (1992),
 7    the experimental variability (indicated by the error bars) on the 2000 ppm data was much larger
 8    than the 200 or 1200 ppm data, and as indicated by the simulations in Figure B-ll  here, there is
 9    at least the appearance that the exposure was actually for 7 hr instead of 6 hr. (To be clear, the
10    2000 ppm data were used in the optimization with the duration of inhalation set to  6 hr, but the
11    routine selected parameters which only poorly fit those data.) Since our greatest concern is in
12    predicting dosimetry at lower exposure levels, near to the points of departure, we decided to
13    retain the fits shown here. The corresponding parameters are listed in Table B-l.  The fractional
14    absorption (20%) was lower than that estimated for mice (66.5%), but Perkins et al. (1995) also
15    found lower fractional absorption of inhaled methanol in rats vs.  mice.
16          Finally, first-order oral absorption parameters were first fit to the lower dose (100 mg/kg)
17    oral absorption data reported by Ward et al. (1997), using the optimization routines in
18    acslXtreme v2.3 (Figure B-12, heavy/solid lines). (Since the animals used were SD rats,  the SD-
19    specific metabolic parameters were used.) While the fit to the low-concentration data was quite
20    good (Figure B-12, lower panel), the fit to the 2500 mg/kg data (Figure  B-12, upper panel)
21    exhibited a much faster and higher peak than shown by the data.  Even when the model was fit to
22    both the high- and low-concentration data simultaneously, the fit to the high-concentration data
23    could not be significantly improved without completely degrading the low-concentration fit (not
24    shown).  Also note that the  2500 mg/kg linear simulation completely over-estimates all the data
25    points; i.e., the area-under-the-curve for this dose is higher than indicated by the data, indicating
26    that the assumption of 100% absorption is not valid.  Therefore, an alternative model using a
27    saturable (Michaelis-Menten) equation for absorption from the stomach and fecal elimination
28    (linear term) from the intestine was considered (thin lines) and found to significantly improve the
29    high-concentration simulation, with  a nearly identical fit the low-concentration data.  While
30    methanol absorption is not known to be regulated by transporters or other processes that would
31    give rise to rate saturation, it is clear form the discrepancy between the linear model and the 2500
32    mg/kg data that uptake is slower than predicted by such a model and its use would lead to an
33    over-prediction of internal concentrations. Therefore parameters for the saturable uptake model
34    are reported in Table B-l and the KMASC applied to mice and humans. Note that since  the
35    saturation constant corresponds to a fairly large dose (620 mg/kg), the model is still effectively
36    linear at low- to moderate dose rates.
                                               B-21         DRAFT - DO NOT CITE OR QUOTE

-------
            E
            3E
2,700

2,400 -

2,100-
1,800

1,500-

1,200-

  900-
  600-

  300-

    0-
                                    2500 mg/kg linear uptake simulation
                                    2500 mg/kg saturable uptake simulation
                                    2500 mg/kg data
                                    100 mg/kg linear uptake simulation
                                    100 mg/kg saturable uptake simulation
                                    100 mg/kg data
0
                                         12
                                    18
                                Time (hr)
                                              24
30
36
            o
            E
            ••_••

            O
            Qj
100 H


 80


 60


 40
            3   20
                                             — 100 mg/kg linear uptake simulation
                                             	 100 mg/kg saturable uptake simulation
                                             n 100 mg/kg data
                   012345678
                                              Time (hr)

         Figure B-12. Model fits to data sets from oral exposures to 100 (squares) or 2,500
1        (diamonds) mg/kg MeOH in female Sprague-Dawley rat (Expanded scale in lower
2        panel). Symbols are concentrations obtained from the command file. The thick lines
3        represent PBPK model fits using a linear (first-order) equation for absorption from
4        the stomach compartment with no fecal elimination, while the thin lines use a
5        Michaelis-Menton equation with a small fraction eliminated in the feces. All other
6        Gl rates, including absorption from the intestine, are first order.

           Source: Ward etal.( 1997).
                                              B-22
                                           DRAFT - DO NOT CITE OR QUOTE

-------
      B..2.7. Rat Model Simulations
 1          A range of adverse developmental effects was noted in rat pups exposed to methanol
 2    throughout embryogenesis (NEDO, 1987).  SD rats were exposed in utero over different periods
 3    of pregnancy and as neonates via inhalation or in drinking water. Inhalation exposures to
 4    methanol were carried out for 18-22 hours, depending on the exposure group. Simulations of
 5    predicted Cmax, AUC, and total metabolized from 22-hour exposures to 500, 1,000, and 5,000
 6    ppmMeOHare shown in Figure B-13. Simulations of oral  exposures of SD rats to 65.9, 624.1,
 7    or 2,177 mg/kg-day (500, 5,000, 20,000 ppm in drinking water), daily dose estimations from the
 8    study of Soffritti et al. (2002), based on measured water consumption, kindly provided by
 9    Cynthia Van Landingham, Environ International, Ruston, Louisiana, are shown in Figure B-13.
10    Although the exposures in these studies are to rats over long periods and in some  cases exposures
11    of the newborn pups, the model simulations are to NP adult rats only, using the dose-group
12    specific average body weights of 0.33-0.34 kg BW from the study of Soffritti et al. (2002) and do
13    not take into account changes is body weight or composition.  These simulated values are
14    presumed to be a better surrogate for and predictor of target-tissue concentrations in developing
15    rats, and the corresponding estimated human concentrations a better predictor of developmental
16    risk in humans than would be obtained using the applied  concentration or dose and default
17    extrapolations. The logic here is simply that the ratio of actual target tissue concentration (in the
18    developing rat pup or human) to the simulated concentration in the NP adult is expected to be the
19    same in both species and hence, that proportionality drops out in calculating a HEC.
20          Figure B-13 depicts simulations run to determine internal doses for 22 hours/day
21    inhalation exposures at 500,  1,000, or 2,000 ppm.  Simulation results for continuous inhalation
22    exposures are  shown for contrast.  The simulations show that for all but the highest dose (2,000
23    ppm) steady state is reached  within 22 hours, and that "periodicity," where the concentration
24    time course is the same for each subsequent day, is reached by the 2nd day of exposure.  At
25    2,000 ppm,  however, steady  state is not reached until  after 48 hours for the continous exposure.
26    Therefore, the Cmax, 24-hour AUC and amount metabolized per day (AMET) were by simulating
27    22 hours/day exposures for 5 days and calculating values of AUC and AMET over the last day
28    (24 hours) of that period.
29
                                               B-23         DRAFT - DO NOT CITE OR QUOTE

-------
          CD
                        12       24      36       48       60
                                              Time (hr)
         72
84
96
Exposure
concentration
(ppm)
500
1,000
2000
^max
(mg/L)
3.6
10.6
48.5
AUC
(hrmg/L
)
79
227
968
AMET
(mgEq)
17.6
34.8
67.2
         Figure B-13.  Simulated Sprague-Dawley rat inhalation exposures to 500,1,000, or
1        2,000 ppm MeOH. Rat BW was set to 0.33 kg.  Simulations are shown for both
2        continuous (thin, dashed/dotted lines in plot) and 22 hours/day exposures (thick, solid lines
3        in plot). Cmax, AUC, and amount metabolized (AMET) are determined from the 22
4        hour/day simulations, run for a total of 5 days (120 hours), with the AUC and AMET
5        calculated for the last 24 hours of the simulation.
1
2          Figure B-14 depicts simulations run to mimic a single oral exposure, treated as a
3   continuous infusion for 12 hours (assuming 12-hour period when rats are awake and active).
4   Total AUC and AMET and AUC24 and AMET24 for the first 24 hours after start of exposure
5   were calculated.
                                             B-24
DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
 4
 5
 6
 7
 8
 9
10
11
12
13
14
15
16
                 10,000-a

                  1,000-
             G~
             "Si    100-
             X
             o
              01
              o
             _o
             CD
               10-

                1

               0.1-J

              0.01
                  0.001
                       0
                                               40
                                            Time (hr)
60
80
Exposure
dose
(mg/kg-day)
66.0
624.1
2177
Body
weight (kg)
0.33
0.33
0.34
^max
(mg/L)
9.3
631.6
2,832.4
AUC
(hrmg/L)
104.8
9,525
72,617
AUC24
(hrmg/L)
104.8
8,817
45,138
AMET
(mgEq)
21.2
155.6
347.4
AMET24
(mgEq)
21.2
122.6
138.4
    Figure B-14.  Simulated rat oral exposures of Sprague-Dawley rats to 65.9, 624.2, or
    2,177 mg MeOH/kg/day. Dosing was simulated as a 12-hour, zero-order infusion to the
    liver compartment.  The AUC and total amount metabolized are given for a period
    sufficient for the MeOH to clear (84 hours), and the AUC24 and AMET24 values represent
    just the first 24 hours of exposure. (Results shown for illustrative purposes. Dosimetry used
    in assessment was simulated using a more realistic water ingestion pattern.)

       To simulate ingestion of methanol in drinking water by rats under bioassay conditions, an
ingestion pattern based on the observations of Spiteri (1982) and Peng et al. (1990).  While mice
ingest water in frequent,  small bouts (Gannon et al., 1992) that are reasonably described as a
continuous delivery to the stomach, rats exhibit clear periods of ingestion alternating with
periods where no ingestion occurs (Spiteri, 1982; Peng et al., 1990).  Based on those data a
reasonable representation of rat water ingestion can be described as serious of pulses. During the
dark/active period of each day (first 12 hr) each bout of drinking was assumed to last 45 min
followed by 45 min without ingestion (total of 8 bouts). During the light/inactive period (next 12
hr) drinking bouts were assumed to last only 30 min followed by 2.5 hr (150 min) without
drinking (4 bouts). An equal amount was assumed to be consumed in each bout within the dark
period, likewise within each light-period bout, with the respective amounts adjusted such that
80% of the total ingestion occurs during the dark and 20% during the light (Burwell et al., 1992).
The resulting  absorption  pattern is shown in Figure B-15A and a simulated blood concentration
time-curve  (for 50 mg/kg/day dosing) is shown in Figure B-15B.
                                               B-25
                                                      DRAFT - DO NOT CITE OR QUOTE

-------
 0,12-
I i
10
*- 0,1-
c
B 0,08-
Q.
0 0,06-
_Q
to 0,04-
C 0,02-
o
4-J
f 1












	











—











—



































	

















A

























-









-









                                             10   12   14   16
                                              Time (hr)
                        18    20   22
                24
                                                                             B
                                  12
18      24     30
    Time  (hr)
36
42
48
         Figure B-15. Rat daily drinking water ingestion pattern (A) and resulting predicted
         blood concentration for a 2-day exposure (B).Rat drinking water exposures were
         simulated by multiplying the fractional absorption rate (1/hr) as a function of clock
         time by the daily total dose ingested (mg) to obtain a rate of addition of methanol
         into the stomach lumen compartment (mg/h).
     B.2.8. Human Model Calibration
      B. 2.8.1. Inhalation Route
 4          The mouse model was scaled to human body weight (70 kg or study-specific average),
 5   using human tissue compartment volumes and blood flows, and calibrated to fit the human
 6   inhalation-exposure data available from the open literature, which comprised data from four
 7   publications (Ernstgard et al., 2005; Batterman et al., 1998; Osterloh et al., 1996; Sedivec et al.,
 8   1981).
 9          A first-order rate of loss of MeOH from the blood, K1C, and a first-order bladder
10   compartment time constant, KBL,  were used to provide an estimate of urinary MeOH
                                             B-26         DRAFT - DO NOT CITE OR QUOTE

-------
 1    elimination. The inhalation-route urinary MeOH kinetic data described by Sedivec et al. (1981)
 2    (Figure B-16) were used to inform these parameters.  The urine MeOH concentration data
 3    reported by the authors were converted to amount in urine by assuming 0.5 mL/hr/kg total
 4    urinary output (Horton et al., 1992). Sedivec et al. (1981) measured a fractional uptake of
 5    57.7%, based on total amount inhaled. Since the PBPK model uses alveolar rather than total
 6    ventilation and this is typically assumed to be 2/3  of total ventilation the fractional uptake of
 7    Sedivec et al. (1981) was corrected by dividing by 2/3 to obtain a value for FRACIN of 0.8655.
 8    The resulting values  of K1C and KBL, shown in Table B-l, differ somewhat depending on
 9    whether first-order or saturable liver metabolism is used.  These are only calibrated against a
10    small data set and  should be considered an estimate. Urine is a minor route of MeOH clearance
11    with little impact on  blood MeOH kinetics.
12           Although the high-doses used in the mouse studies warrant the use of a second metabolic
13    pathway with a high  Km, the human exposure data all represent lower concentrations and may
14    not require or allow for accurate calibration of a second metabolic pathway. Horton et al. (1992)
15    employed two sets of metabolic rate constants to describe human  MeOH disposition, similar to
16    the description used for rats and mice, but in vitro studies using monkey tissues with non-MeOH
17    substrates were used as justification for this approach. Although Bouchard et al. (2001)
18    described their metabolism using Michaelis-Menten metabolism,  Starr and Festa (2003) reduced
19    that to an effective first-order equation and showed adequate fits.  Perkins et al. (1995) estimated
20    a Km of 320 +  1273  mg/L (mean + S.E.) by fitting a one-compartment model to data from a
21    single oral poisoning to an estimated dose. In addition to the extremely high standard error, the
22    large standard error for the associated Vmax (93 + 87 mg/kg/hr) indicates that the set of
23    Michaelis-Menten constants was not uniquely identifiable using this data. Other Michaelis-
24    Menten constants that have been used to describe MeOH metabolism in various models for
25    primates are given in Table B-2.  Because the Km calculated by Perkins et al.  (1995) from the
26    high-dose oral exposure is 320 mg/L, while the highest observed concentration in the data sets
27    considered here is  14 mg/L (Batterman et al., 1998), forcing the model to use this higher Km
28    would simply result in fits that are effectively indistinguishable from the linear model.  A simple,
29    linear model is preferred over the use of a Km value that high.
                                               B-27         DRAFT - DO NOT CITE OR QUOTE

-------
                       10

                       9 -

                       8 -
                    O)
                       3 -

                       2
                       1 -
Sedivec et al., urine
   concentration
	231 ppm saturable sim
	157 ppm saturable sim
	78 ppm saturable sim
	231 ppm linear sim
	157 linear sim
	78 ppm linear sim
 •  231 ppm data
 A  157 ppm data
 O  78 ppm data
                                                                  20
                                                24
                       2.5 -
                               Sedivec et al.,
                           total urinary excretion
                                                                 20
                                                24
         Figure B-16. Urinary MeOH elimination concentration (upper panel) and cumulative
1        amount (lower panel), following inhalation exposures to MeOH in human volunteers.
2        Data points in lower panel represent estimated total urinary MeOH elimination from
3        humans exposed to 78 (diamonds), 157 (triangles), and 231 (circles) ppm MeOH for 8
4        hours, and lines represent PBPK model simulations. Solid lines are model results with the
5        saturable equation for hepatic metabolism while dashed lines show results for liner
6        metabolism. Data digitized from Sedivec et al. (1981) and provided for modeling by the
7        EPA.
           Source: Sedivec et al. (1981).
                                              B-28
                                 DRAFT - DO NOT CITE OR QUOTE

-------
          Table B-2. Primate kms reported in the literature
Km(mg/L)
320 +1273
716 + 489
278
252+116
33.9
0.66
23.7 + 8.7"
Reference
Perkins et al, 1995
Perkins et al, 1995
Perkins et al., 1995
Perkins et al., 1995
Hortonetal., 1992
Fisher etal., 2000
(This analysis.)
Note
Human: oral poisoning, estimated dose
Cynomolgus monkey: 2 g/kg dose
Rhesus monkey: 0.05-1 mg/kg dose
Cynomolgus monkey: 1 g/kg dose
PBPK model: adapted from rat Km
PBPK model, Cynomolgus monkey: 10-900 ppm
PBPK model, human: 100-800 ppm
     Note- the values from Perkins et al. (1995b), are + S.E.
     aMean + S.D. This Km was optimized while also varying Vmax, K1C, and KBL, from all of the at-rest human
     inhalation data as a part of this project. The S.D. given for this analysis is based on the Optimize function of
     acslXtreme, which assumes all data points are discrete and not from sets of data obtained over time and
     therefore a true S.D. would be a higher value. The final value reported in Table B-l (21  mg/L) was obtained by
     sequentially rounding and fixing these parameters, then re-optimizing the remaining ones.  For more detail, see
     text and Table B-3.
 1           To estimate both the Michaelis-Menten and first-order rates, all human data under
 2   nonworking conditions (Batterman et al., 1998; Osterloh et al., 1996; Sedivec et al., 1981) were
 3   used.  Before discussing the parameter estimation, however, adjustments were made to one of
 4   these data sets (Osterloh et al.,  1996). Batterman et al., (1998) and Sedivec et al. (1981) both
 5   subtracted background levels before reporting their results. However, Osterloh et al.  (1996)
 6   measured and reported (plotted) blood methanol in nonexposed controls (data shown in
 7   Figure B-17). The data for Osterloh et al. (1996) clearly show a time-dependent trend which is
 8   close to linear, and a linear regression is also included. However, the blood concentration
 9   (average) in the exposed group of that study was -1.2 mg/L, whereas the  data and regression in
10   Figure B-17 indicate a value of- 0.9 mg/L. Therefore, the exposure data for Osterloh et al
11   (1996) were corrected by  subtracting time-zero value for the exposed group plus  a time-
12   dependent factor obtained by multiplying the slope of this regression (0.093 mg/L-hr) by the
13   measurement time.
14           The metabolic (first-order or saturable) and urinary elimination constants were
15   numerically fit to the nonworking human data sets while holding the value for FRACIN at
16   0.8655 (estimated from the results of Sedivec et al. as described above) and holding the
17   ventilation rate constant at 16.5 L/hr/kg0'75 and QPC at 24 L/hr/kg0'75 (values used by EPA
18   [2000d] for modeling the  inhalation-route kinetics vinyl chloride).  Other human-specific
19   physiological parameters were  set as reported in Table B-l.
                                                B-29
DRAFT - DO NOT CITE OR QUOTE

-------
  1.8
  1.6
J 1.4
oi
5.1.2

I   1
| 0.8
§ 0.6
<§ 0.4
  0.2
    0
                               Control blood concentrations
                                   (Osterlohetal., 1996) •
                                                      y = 0.093x + 0.9094
                                                         R2 = 0.9391
                                                 Time (hr)
          Figure B-17. Control (nonexposed) blood methanol concentrations
             Source: Ernstgard et al. (2005); Osterloh et al. (1996).

 1           Either (a) the set of VmaxC, Km, K1C, and KBL were simultaneously varied while fitting
 2    the entire data set or (b) KLLC, K1C, and KBL were so varied and fit.  Thus the two model fits
 3    are separated by a single degree of freedom (one additional parameter in case [a]).  Statistical
 4    results given in Tables B-2 and B-3 are from these global fitting exercises. Final fitted
 5    parameters that have been used in the model for the risk assessment are given in Table B-l. The
 6    resulting fits of the two parameterizations (1st order or optimized Km/Vmax) are shown in
 7    Figures B-l6 and B-l8.
 8           Use of a first-order rate has the advantage of resulting in one fewer variable in the model
 9    and results in an adequate fit to the data, but the saturable model clearly fits some of the data
10    better (Figures B-l6 and B-l8). To discriminate the goodness-of-fit resulting from the inclusion
11    of an additional variable necessary to describe saturable metabolism versus using a single first-
12    order rate, a likelihood ratio test was performed. Models are considered to be nested when the
13    basic model structures are identical except for the addition of complexity, such as the added
14    metabolic rate. Under these conditions, the likelihood ratio can be used to statistically compare
15    the relative ability of the two different metabolism scenarios to describe the same data, as
16    described by Collins et al. (1999). The hypothesis that one metabolic description is better than
17    another is calculated using the likelihood functions evaluated at the maximum likelihood
18    estimates. Since  the parameters are optimized in the model using the maximum LLF, the
19    resultant LLF is used for the statistical comparison of the models. The equation states that two
20    times the log of the likelihood ratio follows a %2 distribution with r degrees of freedom:

21                  -2[log(/l(modell)//l(model2))] =-2[log/l(modell)-log/l(model2)] =  X2r
                                                B-30
                                      DRAFT - DO NOT CITE OR QUOTE

-------
1    The likelihood ratio test states that if twice the difference between the maximum LLF of the two
2    different descriptions of metabolism is greater than the ^ distribution, then the model fit has
3    been improved (Devore, 1995; Collins et al., 1999; Steiner et al.,  1990).
                                   Batterman et al., blood concentration
                                          (800 ppm exposures)
                                                          -Saturable metabolsim
                                                           First-order metabolism
                                                           2 h data
                                                           1 h data
                                                           30 min data
                                                           Osterloh et al.
                                                        (200 ppm inhalation)
                                     O  Data
                                        Saturable metabolsim
                                        First-order metabolism
                                               Time (hr)
         Figure B-18. Data showing the visual quality of the fit using optimized first-order or
4        Michaelis-Menten kinetics to describe the metabolism of MeOH in humans. The rate
5        constants used for each simulation are given in Table B-3.
           Source: Batterman et al. (1998: top); Osterloh et al. (1996: bottom).

         Table B-3. Parameter estimate results obtained using acslXtreme to fit all human data
         using either saturable or first-order metabolism
Parameters
Michaelis-Menten (optimized)
Km
Optimized value

T3 O
23.8
S.D.

8.8
Correlation matrix
-0.994

LLF
-24.1

                                              B-31
DRAFT - DO NOT CITE OR QUOTE

-------
vmaxc
First Order
KLLC
33.2

95.7
10.1

5.4

NA


-31.0

     Note: The S.D.s are based on the Optimize function of acslXtreme, which assumes all data points are discrete
     and not from sets of data obtained over time and therefore a true S.D. would be a higher value.

 1           At greater than a 99.95% confidence level, using 2 metabolic rate constants (Km and
 2   VmaxC) is preferred over utilizing a single rate constant (Table B-4). While the correlation
 3   coefficients (Table B-3) indicate that Vmax and Km are highly correlated, that is not unexpected,
 4   and the S.D.s (Table B-3) indicate that each is reasonably bounded. If the data were
 5   indistinguishable from a linear system, Km in particular would not be so bounded from above,
 6   since the Michaels-Menten model becomes indistinguishable from a linear model as VmaxC and
 7   Km tend to infinity.  Moreover, the internal dose candidate POD, for example the BMDLio for
 8   the inhalation-induced brain-weight changes from NEDO (1987), with methanol blood AUC as
 9   the metric, is 374.67 mg-hr/L, which corresponds to an average blood concentration of 15.6
10   mg/L. Therefore the Michaelis-Menten metabolism rate equation appears to be sufficiently
11   supported by the existing  data, and its use is expected to improve the accuracy of the HEC
12   calculations, since those are being conducted in a concentration range in which the nonlinearity
13   has an impact.

          Table B-4. Comparison of LLF for Michaelis-Menten and first-order metabolism
LLF(logI)forM-M
-24.1
LLF (logl) for 1st
order
-31.0
LLF
1st versus M-Ma
34.1
xl (99% confidence)13
13.8
X2r (99.95%
confidence) b
12.22
      Note: The models were optimized for all of the human data sets under non working conditions. M-M:
      Michaelis-Menten
      "obtained using this equation: - 2[log /l(model l) - log /l(model 2)]
      ^significance level at r = 1 degree of freedom.

14           While the use of Michaelis-Menten kinetics might allow predictions across a wide
15    exposure range (into the nonlinear region),  extrapolation above 1,000 ppm is not suggested since
16    the highest human exposure data are for 800 ppm.  Extrapolations to higher concentrations are
17    potentially misleading since the nonlinearity in the exposure-internal-dose relationship for
18    humans is uncertain above this point. The use of a BMD or internally applied UFs should place
19    the exposure concentrations within the range of the model.
20           The data from Ernstgard et al. (2005) was used to assess the use of the first-order
21    metabolic rate constant to a dataset collected under conditions of light work. Historical measures
22    of QPC (52.6 L/hr/kg0'75) and QCC (26 L/hr/kg0'75) for individuals exposed under conditions of
23    50 w of work from that laboratory (52.6 L/hr/kg0'7) (Ernstgard, personal communication; Corley
                                                B-32
DRAFT - DO NOT CITE OR QUOTE

-------
 1   et al., 1994; Johanson et al., 1986) were used for the 2-hour exposure period (Figure B-19).
 2   Otherwise, there were no changes in the model parameters (no fitting to these data).  The results
 3   are remarkably good, given the lack of parameter adjustment to data collected in a different
 4   laboratory, using different human subjects than those to which the model was calibrated.
                                        Ernstgard et al. (2005)
 5
 6
 7
 8
 9
10
 5
 6
 7
 8
 9
10
11
12
13
14
15
16
17
18
                      OQ
                                              4         6
                                                Time (hr)
                                                                     10
    Figure B-19. Inhalation exposures to MeOH in human volunteers. Data points
    represent measured blood MeOH concentrations from humans (4 males and 4 females)
    exposed to 100 ppm (open symbols) or 200 ppm (filled symbols) for 2 hours during light
    physical activity.  Solid lines represent PBPK model simulations with no fitting of model
    parameters. For the first 2 hours, a QPC of 52.6 L/hr/kgO.75 (Johansen et al. 1986), and a
    QCC of 26 L/hr/kgO.75 (Corley et al., 1994) were used by the model.
       Source:  Ernstgard et al.(2005).

 B.2.8.1. (Should be B.2.8.2) Oral Route
       There were no human data available for calibration or validation of the oral route for the
human model. In the absence of data to estimate rate constants for oral uptake, the 'humanset.m'
file which sets parameters for human simulations applies the KMAS for the mouse with the other
absorption parameters set to match those identified for ethanol in humans by Sultatos et al.
(2004); VmASC was set such that for a 70-kg person, VMAS/KMAS matches the first-order
uptake constant of Sultatos et al. (2004) (0.21 hr"1). While Sultatos et al. include a term for
ethanol metabolism in the stomach, no such term is included here and the rate of fecal
elimination is set to zero, corresponding to 100% absorption. However zero-order ingestion, a
continuous infusion at a constant rate into the stomach lumen equal to the daily dose/24 hours,
was assumed for all human simulations. Since absorption was assumed to be  100% of
administered MeOH, at steady state the rate of uptake from the stomach and intestine
compartments (combined) must equal the rate of infusion to the stomach. Since Cmax is driven by
the oral absorption rate, which was assumed rather than fitted and verified, Cmax was not used as
a dose metric for human oral route simulations. AUC, which is less dependent on rate of uptake,
                                              B-33
                                                      DRAFT - DO NOT CITE OR QUOTE

-------
 1   was used as the dose metric and for estimation of HEDs.  Since the AUC was computed for a
 2   continuous oral exposure, its value is just 24-hours times the steady-state blood concentration at
 3   a given oral uptake rate.

      B. 2.8.1. (B. 2.8.3) Inhalation Route HECs and Oral Route HEDs
 4          The atmospheric MeOH concentration resulting in a human daily average blood MeOH
 5   AUC (hrxmg/L) or Cmax (mg/L) equal to that occurring in experimental animals following
 6   exposure at the POD concentration is termed the HEC. Similarly, the oral dose (rate) resulting in
 7   human daily average blood MeOH AUC (hrxmg/L) equivalent to that occurring in an
 8   experimental animal at the POD concentration is termed the HED.
 9          To determine the HEC for specific exposures in mice, the mouse PBPK model is first
10   used to determine the daily blood MeOH 24-hours AUC and Cmax associated with 7 hour/day
11   inhalation exposures. Mice were exposed each day for 10 days, so the full 10-day exposure was
12   simulated and an average 24-hours AUC calculated over that time, so no other duration
13   adjustment was needed. The human AUC was determined for the last 24 hours of a continuous
14   1,000-hour exposure, to assure steady state was achieved. The human Cmax was determined at
15   steady state and so is equivalent to the steady state blood MeOH concentration. Results are given
16   in Table B-5 and for inhalation shown in Figure B-20.
17          For example, for a 1,000 ppm exposure this  resulted in model-predicted peak blood of
18   133 mg/L and an AUC of 770 (hrxmg/L). The human model can then be used to determine the
19   human MeOH exposure concentration leading to the same daily average AUC or Cmax under
20   continuous exposure conditions. Based on AUC, the HEC of the 1,000-ppm exposure is
21   684 ppm, while based on peak (human steady-state) concentration, the HEC is predicted to be
22   1110 ppm. The parameters used in the human model for these simulations are listed in Table B-l
23   for saturable kinetics.
24          The HED was calculated by using the human model to find the oral dose (mg/kg-day) that
25   gave a blood MeOH AUC equivalent to the mouse AUC following an exposure at the POD.
26   Zero-order absorption was assumed. For example, the human oral exposure equivalent to a
27   1,000-ppm inhalation exposure in mice (i.e., with an AUC of 770 mg-hr/L) is 163 mg/kg-day.
28   Since a 200 mg/kg-day oral exposure gave a human AUC of 1,320 mg-hr/L, which falls between
29   the values predicted for inhalation exposures at 800 ppm (1,090 mg-hr/L) and 1,000 ppm (2,090
30   mg-hr/L), this oral exposure rate was taken to be the upper end for the model to accurately
31   estimate an HED.
                                             B-34        DRAFT - DO NOT CITE OR QUOTE

-------
         Table B-5. PBPK model predicted Cn
         exposed to MeOH
and 24-hour AUC for mice and humans
Exposure
concentration
(ppm)
1
10
50
100
250
500
1,000
2000
5000
1,0000

Inhalation Route
Mouse"
AUC
(mg-hr/L)
0.15
1.52
7.98
17.0
53.4
170
770
3310
17300
51200
Cmax (mg/L)
0.021
0.22
1.14
2.45
7.77
26.1
133
524
2000
4710

Humanb
AUC
(mg-hr/L)
0.59
5.97
30.6
63.3
177
447
2090
31100
147000
341,000
Css (mg/L)
0.025
0.25
1.28
2.64
7.36
18.6
87.2
1300
6130
14200
Oral Route
Human
Dose
(mg/kg-day)
0.1
1
10
50
100
200
500
1,000
2000
5000
AUC
(mg-hr/L)
0.204
2.05
21.2
124
315
1320
39400
125000
297000
814000
     "The mouse 24-hour average AUC were calculated under the conditions of the bioassay: 10 days of exposure
     with 7 hours of exposure during each 24 hour period.
     bHuman simulation results are considered unreliable above 1,000 ppm (inhalation) or 200 mg/kgday (oral), but are
     included for comparison

1          Again, since the available human exposure data is to, at most, 800 ppm, the model could
2    not be calibrated for higher exposures that approximate most of the mouse and rat exposure
3    concentrations. The AUC in humans for similar exposure levels is ~3 times greater than in the
4    mouse, primarily because human exposure estimates are expected to result from 24-hour
5    exposures and the mice were exposed for 7 hours.
                                               B-35
            DRAFT - DO NOT CITE OR QUOTE

-------
          1000000
           100000 -
            10000 -
•?
O)
E.
O
<
             1000 -
              100 -
               10 -
                1 -
              0.1
                          •Mouse
           	Human
           	Human-linear
                  1      10    100   1000  10000
                       Concentration (ppm)
                                          10000
                                                      10     100    1000   10000
                                                     Concentration (ppm)
                            10000
                         "8  1000

                         I iT100 H
                            °
                 o
                 <
i 10 -

  1 -

 0.1
                                                  	Human
                                                  	Human-linear
                                          10       100      1000
                                          Concentration (ppm)
                                                         10000
          Figure B-20. Predicted 24-hour AUC (upper left), Cmax (upper right), and amount
          metabolized (lower) for MeOH inhalation exposures in the mouse (average over a 10-
          day exposure at 7 hours/day) and humans (steady-state values for a continuous
          exposure). Cmax for human exposures is equal to the steady-state blood concentration. For
          humans, the long-dashed lines are model predictions using Michaelis-Menten
          metabolism (optimized Km of 23.8 mg/L) and the short-dashed lines are model
          predictions using first-order kinetics.  Amount metabolized normalized to BW° 75 to
          reflect cross-species scaling (Human simulation results above 1,000 ppm are not
          considered reliable but are shown for comparison).

 1          While the PBPK computational code can be used in the future to derive HECs or HEDs
 2   for other exposures, an alternative approach was developed that allows non-PBPK model users to
 3   estimate MeOH HECs and HEDs from benchmark doses in the form of AUCs.  This approach
 4   uses algebraic equations describing the relationship between predicted MeOH 24-hour AUC or
 5   total amount metabolized in the liver (MET, mg [per day]) (constant 24-hour exposure) and the
 6   inhalation exposure level (i.e., an HEC in ppm) (Equations  1, Ib, 3 or 3b) or oral exposure rate
 7   (i.e., an HED in mg/kg-day) (Equations 2, 2b, 4 or 4b). To use the equations to derive an HEC or
 8   HED, the target human AUC is applied to the appropriate equation. Since these relationships are
 9   for continuous exposures, blood concentration is constant, and hence extrapolation for a Cmax is
10   obtained by simply using AUC = 24*Cmax.
                                             B-36
                                                 DRAFT - DO NOT CITE OR QUOTE

-------
 4

 5

 6

 7
JTLC,^ ^PJJIH
Rf C (ppm)
HED (mg/kg/
RfD (mg/kg/d£
HEC(ppm
-------
 1    species, and hence that using NP maternal blood levels in place of fetal concentrations will not
 2    lead to a systematic error when extrapolating risks. Thus, a full representation of pregnancy and
 3    the fetal/conceptus compartment appears to be unnecessary.
 4           While lactational exposure is less direct than fetal exposure and blood or target-tissue
 5    levels in the breast-feeding infant or pup are likely to differ more from maternal levels, the
 6    health-effects data indicate that most of the effects of concern are  due to fetal exposure, with
 7    only a small influence due to postbirth exposures.  Separating out the contribution of postbirth
 8    exposure from pre-birth exposure to a given endpoint in a way that would allow the risk to be
 9    estimated from estimates of both exposure levels would be extremely difficult, even if one had a
10    lactation/child PBPK model that allowed for prediction of blood (or target-tissue) levels in the
11    offspring. And one would still expect the target-tissue  concentrations in the offspring to be
12    closely related to maternal blood levels (which depend on ambient exposure and determine the
13    amount delivered through breast milk), with the relationship between maternal levels and those
14    in the offspring being similar across species.
15           Therefore, the development  of a lactation/child PBPK model appears not to be supported,
16    given the minimal change that is likely to result in risk extrapolations and use of (NP) maternal
17    blood levels as a measure of risk in the offspring is still considered preferable over use of default
18    extrapolation methods.  In particular, the existing human data allow for accurate predictions of
19    maternal blood levels, which depend strongly on the rate of maternal methanol clearance.
20    Failing to use the existing data (via PBPK modeling) on human methanol clearance (versus that
21    in other species) would be to ignore this very important determinant of exposure to breast-fed
22    infants. And since bottle-fed infants do not receive methanol from their mothers, they are
23    expected to have lower or, at most, similar overall exposures for a given ambient concentration
24    than the breast-fed infant, so that use of maternal blood levels for  risk estimation should also be
25    adequately protective for that group.
26           During model development, several inconsistencies between experimental blood MeOH
27    kinetic data embedded in the Ward et al. model (1995)  and the published figures first reporting
28    these data were discovered.  Therefore, data were digitized from the published literature when a
29    figure was available, and the digitized data was compared to the provided data. When the
30    digitized data and the data embedded in the computational files (i.e., provided to Battelle under
31    contract from the EPA) were within 3% of each other, the provided data was used; when the
32    difference was greater than 3%, the  digitized data was used.  Often,  using the published figures
33    as a data source resulted in substantial improvements of the fit to the data in the cases where the
34    published figures were different from the embedded data.
35           The final MeOH PBPK model fits well inhalation-route blood kinetic data from separate
36    laboratories in rodents and humans. Intravenous-route blood MeOH kinetic data in NP mice
37    were only available for a single i.v. dose of 2,500 mg/kg, but were available for GDIS mice
38    following administration of a broader range of doses: 100, 500, and 2,500 mg/kg. Up to
                                                B-3 8         DRAFT - DO NOT CITE OR QUOTE

-------
 1    20 hours postexposure, blood MeOH kinetics appear similar for NP and pregnant mice after
 2    administration of 2,500 mg/kg. The intravenous pharmacokinetic data in GDIS mice showed an
 3    unexpected dose-dependent nonlinearity in initial blood concentrations, suggesting either a dose
 4    dependence on the volume distribution, which is unlikely, or some source of experimental
 5    variability. To account for this nonlinearity, Ward et al. (1997) used dose-specific partition
 6    coefficients for placenta and embryonic fluid and Vmax for the metabolism of MeOH. The
 7    current model uses a consistent set of parameters that are not varied by dose and therefore does
 8    not fit these 100 mg/kg dose intravenous data. The model does fit the 500 and 2,500 mg/kg
 9    doses, and if a presumed i.v. dose of 200 mg/kg (twice the reported 100 mg/kg) is employed, is
10    able to predict initial blood concentrations for the lowest dose data, as expected. The i.v. data
11    from the Ward et al. (1995) model does match the corresponding published figures.
12          The model fits to the mouse oral-route MeOH kinetic data using a consistent set of
13    parameters (Figure B-4) are reasonably good but not as good as fits to the inhalation data. The
14    model consistently underpredicts the amount of blood MeOH reported in two studies (Ward
15    et al., 1997, 1995). Ward et al. (1997) utilized a different Vmax for each oral absorption data set.
16    In the report by Ward et al. (1997) the GDIS and the GD8 data from Dorman et al. (1995) were
17    both fit using a Vmax of-80 mg/kg/hr (body weights were not listed, the model assumed that
18    GD8 and GDIS mice were both 30 g; Ward et al. (1997) did not scaled by body weight), but
19    lower partition coefficients for placenta (1.63 versus 3.28) and embryonic fluid (0.0037 versus
20    0.77). The current model adequately fits the oral pharmacokinetic data using a single set of
21    parameters that is not varied by dose or source of data.
22          The fits of the rat model to the limited dataset readily available were quite good. The low-
23    dose exposures of all routes were emphasized in model optimization since they were the doses
24    most relevant to risk assessment. Based on a rat inhalation exposure to 500 ppm, the human HEC
25    would be 301 ppm (by applying an AUC of 226  [Figure B-12] to Equation  1).
26          The mouse, rat, and human models fit multiple datasets from multiple research groups
27    using consistent parameters that are representative of each species, but are not varied within
28    species. Using the model, it will be possible to ascertain chronic human exposure concentrations
29    that are likely to be without an appreciable risk of deleterious effects.

      B.3. ADDITIONAL MATERIAL

30       •  Results from Optimizations
31       •  acslXtreme Program (.csl) File (Electronic Attachment)
32       •  acslXtreme procedure (.cmd) file
33       •  Key to .m files for reproducing the results in this report
34       •  Code for .m files

                                              B-39        DRAFT - DO NOT CITE OR QUOTE

-------
 1       •  Personal communication from Lena Ernstgard regarding human exposures reported in the
 2          Ernstgard and Johanson, 2005 SOT poster
 3       •  Personal communication from Dr. Rogers regarding mouse exposures to MeOH
 4       •  Data and simulations for MeOH Metabolic Clearance/Total Metabolites Produced
 5       •  Multiple daily oral dosing for humans

     B.3.1. Results From Optimizations
       B.3.1.1. Approach for and Results of the Optimization of Metabolic Parameters and
       Inhalation Route Fractional Availability in Mice
 6      The approach and results are presented below in outline format with supporting figures. More
 7   complete documentation was not developed because the products of the optimizations were not
 8   used in the final model. The documentation here is intended only to demonstrate that appropriate
 9   optimizations were conducted and what the results of those optimizations were.
10   1.  The Vmax for the low affinity pathway was set to 0  and the remaining VmaxC, Km, and
11       fractional availability were optimized using inhalation data only.
12       a.  The optimizer was unable to find a value for Km that was greater than 0.
13       b.  The resulting metabolic parameters essentially represented a zero order loss process.
14   2.  The Vmax for the low affinity pathway was set to 0 and the remaining VmaxC and Km were
15       optimized using all (oral, intravenous and inhalation) data.
16       a.  The optimized single Km, 135 mg/L, was equal to the average of the 2 original Kms.
17       b.  Fits to the MeOH blood levels following inhalation exposures > 2,000 ppm  are slightly
18          improved, but the model fits  to the  1,000 ppm  exposure concentration  overpredict
19          reported values by 20%.
20   3.  Parameters for both metabolic pathways were optimized using all (oral, intravenous and
21      inhalation) data.
22       a.  The fit  to the high-dose intravenous data from Ward etal. (1997) (2,500  mg/kg) was
23          improved (Figure B-21).
24       b.  The fit  to the  high-dose oral data, also from Ward etal. (1997), (2,500  mg/kg)  was
25          improved (Figure B-22).
26       c.  The fit to the mid-dose i.v. data (500 mg/kg) dose was not as good as using the visually
27          fit parameters (Figure B-21)
28       d.  The fit to the low-dose oral  data (1,500 mg/kg) was not as good when the visually fit
29          parameters were used (Figure B-22). The low-dose data was from Dorman et al. (1995).

                                              B-40        DRAFT - DO NOT CITE OR QUOTE

-------
        e.  Neither set of parameters resulted in an  adequate fit to the low-dose intravenous data
           (100 mg/kg; Figure B-21).

        f.  Fits to the inhalation data following exposures to < 5,000 ppm MeOH were substantially
           worse than when using the visually fit parameters (Figure B-23)
                 10000
                 1000
               o
               
-------
         Figure B-22. Fit of the model to oral data using different clearance and uptake
1        parameter optimizations. Solid blue lines - visually optimized; dashed red lines -
2        clearance parameters (Km, Km2, VmaxC, Vmax2C) optimized using all inhalation data sets;
3        dash/dot green lines - clearance parameters optimized using all data sets (inhalation, oral,
4        and intravenous).
           Source: Ward et al. (1997); Dorman et al. (1995).
         Figure B-23. Fit of the model to inhalation data using different clearance and uptake
1        parameter optimizations. Dotted black lines - model optimized fractional inhalation, solid
2        blue lines - visually optimized; dashed red lines - clearance parameters (Km, Km2, VmaxC,
3        Vmax2C) optimized using all inhalation data sets; dash/dot green lines - clearance
4        parameters optimized using all data sets (inhalation, oral, and intravenous).
           Source:  Rogers et al. (1997).
                                              B-42
DRAFT - DO NOT CITE OR QUOTE

-------
       B. 3.1.1. [B3.1.2.J Conclusion

 1           Under the best circumstances, formal optimizations offer the benefit of repeatability and
 2    confirmation that global optima have not been missed by user-guided visual optimization.
 3    Incorporating judgments regarding the value of specific data sets while easy when visually
 4    fitting,  is difficult at best when using optimization routines. This is an important distinction
 5    between these approaches for this modeling exercise.
 6           The mouse NOEL was 1,000 ppm MeOH. Fitting the blood MeOH concentration data at
 7    this exposure drove our modeling exercises because of the importance of this exposure group in
 8    the risk assessment. Unfortunately, the vast majority of the blood MeOH data came from much
 9    higher exposures. As expected, our various attempts at optimization led to fits that were better
10    for some, but never all,  data sets. This is to be expected when there is clearly significant
11    variability in the underlying data. Various data weighting schemes were included to improve
12    overall optimization while maintaining a good fit to the 1,000 ppm data. In the end, optimization
13    offered no significant improvement over the fractional uptake and metabolic parameter values
14    obtained by visual optimization, so these were retained in the final version of the model.

      B.3.2. acslXtrerne Program (.csl)File

15    PROGRAM MeOH - PBPK Model for Methanol
16    PROGRAM MeOH - PBPK Model for Methanol
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44    INITIAL
45    ! Initialize some Variables before start
Based on MeOH Model by Ward et al 1997 with these revisions:
TS Poet, P Hinderliter and J Teeguarden,
Center for Biological Monitoring and Modeling  4/16/05
Pacific Northwest National Laboratory
Model contains inhalation, iv, and oral (multiple patterns).
1) Removed fetal compartment and other tissues that could be lumped
based on similarity of partition coefficients or did not need to be
specified directly (Bone, mammary tissue) for the modeling purposes here.
2) Changed day to hr.
3) Flows (scaled to BWor BW**0.75), Metabolism (BW**0.75) and
tissue volumes (BW) are scaled in the model.
Final has somach and intestine compartments which provide fast and
slow absorption rates, respectively.
4) Bladder compartment (for human simulations) added by Paul
Schlosser, U.S. EPA, Oct. 2008
5) "Sipping" drinking water exposure code for rats, to match data
from Pengetal. (1990)
6) Time-variable drinking pattern for mice from Keys et al. (2004)
added by Paul Schlosser, U.S. EPA, Aug. 2009
Version is final version used for simulations
     ***** MODEL UNITS***
      Concentration, mg/L
      Mass of Chemical, mg
      Volume, L
      Flow, L/hr
      Body Weight Kg
                            =72 Character Line=
                                                B-43         DRAFT - DO NOT CITE OR QUOTE

-------
 1    Integer IDS, MULIE
 2    REAL DRT(6), DRP(6)  Istore drink water times, percents in array!
 O
 4    CONSTANT   BW = 0.030   ! Body weight (kg)
 5    CONSTANT   QPC = 15.    ! Alveolar ventilation (L/hr/kg**0.75)
 6
 7    ! Blood Flows (fraction of cardiac output)
 8          CONSTANT   QCC=15.0    ! Cardiac output (L/hr/kg**0.75)
 9          CONSTANT   QFC = 0.05    ! Fat
10          CONSTANT   QLC = 0.25    ! Liver
11    ! Blood flow to rest of body Calculated by Flow Balance!
12          QRC= 1.0-(QFC + QLC)
13          QC = QCC*BW**0.75
14          OP = QPC*BW**0.75
15
16    ! Tissue Volumes for mice (fraction of body weight)
17          CONSTANT   VAC = 0.0123  ! Arterial blood
18          CONSTANT   VFC = 0.07    ! Fat
19          CONSTANT   VLC = 0.055   ! Liver
20          CONSTANT   VLuC = 0.0073 ! Lung tissue
21          CONSTANT   VVBC = 0.0368 ! Venous blood
22          VRC = 0.91 - (VAC+VFC+VLC+VLuC+VVBC)
23
24    ! Partition Coefficients (Mouse values from Ward et al used as default)
25          CONSTANT   PB = 1350     ! MeOH Blood:Air; Use Morton value!
26          CONSTANT   PF = 0.08      ! MeOH Fat:Blood
27          CONSTANT   PL = 1.1       ! MeOH Liver:Blood
28          CONSTANT   PLU = 1.0      ! MeOH Lung:Blood, compartment for dosing only
29          CONSTANT   PR = 0.8      ! MeOH Rest of body:Blood
30
31    ! Hepatic Metabolism of MeOH
32          CONSTANT   KM = 45.0     ! mg/L
33          CONSTANT   VMAXC =  15.0 ! mg/hr/BW**0.75
34          VMAX = VMAXC*BW**0.75    ! mg/hr
35          CONSTANT   VMAX2C = 15.0! 2nd saturable pathway
36          VMAX2 = VMAX2C*BW**0.75
37          CONSTANT   KM2= 45.0
38          CONSTANT KLLC = 0.0       ! First-order metabolism
39    ! Set VMAXC = VMAX2C = 0, when KLLC > 0
40          KLL = KLLC/BW**0.25
41
42    ! MeOH Clearance from Blood!
43          CONSTANT   K1C = 0.01    ! First-order clearance, BW**0.25/hr
44          K1 =K1C/BW**0.25   ! Scaled blood elimination, hr-1
45    ! This lumped term was used in the WARD model an accounted for
46    ! renal elimination and "additional" non-hepatic metabolism of
47    ! MeOH associated only with high dose i.v. data.
48    ! A 1st-order term should not be used to represent two processes
49    ! with different dose-dependencies.
50    ! This has not been used for mouse data (set=0), but was  used to
51    ! approximate human urinary data!
52
53    ! Bladder compartment added (Paul Schlosser, U.S. EPA, 10/2008)
54          CONSTANT KBL=0.0  ! Bladder constant, 1/hr
55
56    ! Fractional Absorption of MeOH
57          CONSTANT   FRACin = 0.85 ! Inhalation, value from Perkins et al
                                             B-44        DRAFT - DO NOT CITE OR QUOTE

-------
 1          CONSTANT   KFEC = 0.0    ! Fecal elimination constant, 1/hr
 2                ! KFEC determines oral bioavailability
 O
 4   ! Molecular Weight of MeOH
 5          CONSTANT   MWMe = 32.0  ! mol wt, g/mol!
 6
 7   ! Closed Chamber Parameters
 8          CONSTANT   VChC = 100.0  ! Volume of closed chamber (L)
 9          CONSTANT   Rats = 0.0     ! Number of rats in chamber
10          CONSTANT   kLoss = 0.0    ! Chamber loss rate /hr
11          ! Set RATS = 0.0 and KLOSS = 0.0 for open chamber
12
13   ! Blood Flows (L/hr)
14          OF = QFC*QC ! Fat
15          QL = QLC*QC ! Liver
16          OR = QRC*QC ! Rest of Body
17
18   ! Tissue Volumes (mL)
19          VAB = VAC*BW! Arterial blood volume
20          VF = VFC*BW        ! Fat
21          VL = VLC*BW        ! Liver
22          VLu = VLuC*BW       ! Lung
23          VR = VRC*BW        ! Rest of the body
24          VVB = VVBC*BW      ! Venous blood
25          VBL = VAB + VVB     ! Total blood
26
27   !	Timing commands	!
28          CONSTANT   TCHNG = 6.0        ! End of exposure!
29          CONSTANT   TSTOP = 24.0        ! End of experiment/simulation!
30          CONSTANT   POINTS = 1000.0     ! No. points for simulation output!
31          CONSTANT   REST = 100000.0     ! End of work period for human exercise
32          CONSTANT   WORK = 100000.0    ! Start of work period for human exercise
33          SCHEDULE DS1 .AT.REST    ! Change from work to rest conditions
34          SCHEDULE DS2.AT.WORK   ! Change from rest to work conditions
35          ! Human Rest/Work (changes in blood-flow fractions to fat/liver not currently usedO
36          CONSTANT QPCHR=15.0, QCCHR=15.0,  QLCHR=0.25, QFCHR=0.05 ! Rest
37          CONSTANT QPCHW=52.0, QCCHW=26.0, QLCHW=0.16, QFCHW=0.06      ! Work
38
39   |	Simulation Control	!
40   ! Exposure Conditions Based on User Defined Initial Amounts of
41   ! Chemical (mg)
42          CONSTANT CONCppm = 0.0   ! Air Concentration in ppm
43          VCh  = VChC-(Rats*BW)             ! Volume of Occupied Chamber
44          CONCmg = CONCppm*MWMe/24451  ! Convert ppm to mg/Liter!
45          ACHO = CONCmg*VCH             ! Init Amt in  Chamber, mg!
46
47   ! Oral dosing
48          CONSTANT   KAS = 0.1     ! 1st order oral abs,  hr-1
49          CONSTANT   KMASC = 550  ! Saturable oral abs Kmasc [=] mg/kg
50                KMAS = KMASC*BW
51          CONSTANT   VASC  = 1740 ! Saturable oral ab VmaxC, mg/hr/kgA0.75
52                VAS = VASC*BW**0.75 ! Saturable oral ab Vmax, mg/hr
53          CONSTANT   KAI = 0.1      ! 1st order oral abs from intestine, hr-1
54          CONSTANT   KSI = 0.5     ! 1st order transfer stom to intes hr-1
55          CONSTANT   DOSE = 0.0   ! Oral dose in mg/kg BW
56          CONSTANT   ODS = 0.0     ! Switch for zero order oral uptake
57          ! (Set to 1 for zero order, set to 0 for first order)
                                            B-45        DRAFT - DO NOT CITE OR QUOTE

-------
 1           ODOSE = DOSE*BW*(1.0-ODS)       ! Convert mg/kg to mg total (oral)
 2           RAOZ = DOSE*BW*ODS/24.0  ! mg/hr for zero order dosing
 O
 4           ! Daily dose for steady drinking water by "sipping" (by rats)
 5                 CONSTANT DWDOSE = 0     ! mg/kg/d by periodic sipping
 6                 CONSTANT PER1 = 1.5       ! Period between sipping episodes (hr) during dark
 7           ! "Between" means from the start of 1 to the start of the next episode
 8                 CONSTANT DUR1 = 0.75      ! Duration of sipping episodes during dark (hr)
 9                 CONSTANT PER2 = 3.0       ! Period during light (hr) between sipping episodes
10                 CONSTANT DUR2 = 0.5       ! Duration of sipping episodes during light (hr)
11                 CONSTANT FNIGHT = 0.8     ! Fraction of drinking during night
12                 constant days = 7.0    ! days/week of oral exposure
13                 constant metd = 7.0    ! number of days at end over which AUCBF and AMETF
14                                             ! are averaged
15                 tmetf = metd*24.0
16                 dayon=24.0*days
17           ! Night sipping rate (mg/h) during episodes
18                 DWRNIGHT = DWDOSE*BW*FNIGHT*PER1/(12.0*DUR1)
19           ! Day sipping rate (mg/h)  during episodes
20                 DWRDAY=  DWDOSE*BW*(1-FNIGHT)*PER2/(12.0*DUR2)
21           IDOSE=0
22           ! Above assumes 12-hr each for day/night
23
24    ! Drinking Table from Deborah Keys for mice, as used in
25    ! A quantitative description of suicide inhibition of dichloroacetic acid in rats and mice.
26    ! Keys DA, Schultz IR, Mahle DA, Fisher JW.
27    ! Toxicol Sci. 2004 Dec;82(2):381-93.
28    ! Based on data of Yuan, J. (1993). Modeling blood/plasma concentrations in dosed feed and dosed
29    ! drinking water toxicology studies. Toxicol. Appl. Pharmacol. 119, 131-141.
30    constant rdrink =1.0! Default for use of sipping w/ DWDOSE abopve
31           ! set rdrink = 0.0 to use pattern below
32    table mdrinkp,1,49 / 0., .5, 1., 1.5, 2., 2.5, 3., 3.5, &
33         4., 4.5, 5., 5.5, 6., 6.5, 7., 7.5, &
34         8., 8.5, 9., 9.5, 10., 10.5, 11., 11.5,  &
35        12., 12.5, 13., 13.5, 14., 14.5, 15., 15.5, &
36        16., 16.5, 17., 17.5, 18., 18.5, 19., 19.5, &
37        20., 20.5, 21, 21.5, 22., 22.5, 23., 23.5, 24.0,  &
38         0.12 , 0.9, 1.6, 1.8, 1.9, 2.9, 4.0, 4.5, 4.9, 4.9,  &
39        4.8,4.4,4.0,5.0,5.9,5.3,4.5,3.9, &
40        3.2, 3.0, 2.7, 2.5, 2.3, 2.3, 2.3, 1.9, &
41        1.4, 1.4, 1.3, 1.3, 1.3, 1.1,0.8,0.8, &
42        0.8, 0.6, 0.5, 0.7, 0.8, 0.6, 0.4, 0.2, &
43        0.05, 0.08, 0.14, 0.07, 0.06,  0.08, 0.12 /
44
45           ! Larger bolus dosing
46                 CONSTANT DRDOSE=0.0     ! Total dose by drinking water in boluses, mg/kg day
47           ! Times for multiple oral drinks/day *after* 0
48                 !  Must be ascending, 0 <= times < 24 hr
49                 !  CONSTANT DRT=0, 2,4,6,8,10     ! Rat values
50                 Constant DRT = 0.0, 3.0, 5.0, 8.0, 11.0, 15.0   ! Human values
51                 !  DRTIME(1) assumed = 0 and not used
52           ! Fraction consumed  by drinking at those times
53                 CONSTANT DRP = 0.25, 0.1, 0.25, 0.1, 0.25, 0.05
54
55           !Total oral bolus dose; initial value given at t=0 via initial condition
56                 TODOSE = DRP(1)*DRDOSE*BW*(1.0-ODS) + ODOSE
57
                                                B-46        DRAFT - DO NOT CITE OR QUOTE

-------
 1   ! IV dosing
 2          CONSTANT IVDOSE = 0.0     ! IV dose, mg/kg
 3          CONSTANT TINF = 0.025      ! Length of exposure (hrs), default = 1.5 min (bolus)
 4                ! 1.5 min reported by Ward and Pollack, DMD 1996
 5          TIV = IVDOSE*BW    ! Expected amt infused, mg
 6          IV1 = TIV/TINF ! Rate of infusion, mg/hg
 7
 8   ! For I.V. Runs, control step size if necessary by changing MaxT, not POINTS or CINT
 9          MAXT = 1.0    ! Maximum Step Size, Hours
10          !IF (IVDOSE.GE.1.0E-4) MAXT = 1.0E-4
11
12   ! Liver infusion
13          CONSTANT LIVRO = 0.0       ! Zero-order liver total, mg/kg/day
14          RLIVO = LIVRO*BW/TCHNG    ! Rate in mg/hr
15
16   !	Dose Scheduling	
17          CONSTANT MULTE=0              ! Default is *no* repeated dosing/inhalation
18          CIZON E = 1.0  ! Start with  inhalation on
19          IVZON E = 1.0  ! Start with  IV on
20          SCHEDULE OFF.AT.TCHNG   ! Turn off exposure at TCHNG
21          DAY = 0;
22          NEWDAY = 0; IDS = 2  ! First dose given as initial condition
23          IF (MULTE) SCHEDULE ORALDOSE.AT.DRT(2)
24   ALGORITHM IALG = 2 ! Gear algorithm
25   END   ! END OF INITIAL
26
27   DYNAMIC
28
29   DERIVATIVE
3Q   ,.„„„„„„.„.„„„„  MeOH  .........................
31          IVR = IVZONE*IV1    ! IV dosing; IVR = ate of infusion, mg/hg
32   ! Oral Dosing
33          OWING = ((DWRNIGHT*PULSE(0.0,PER1,DUR1)*PULSE(0.0,24.0,12.0) + &
34                DWRDAY*PULSE(0.0,PER2,DUR2)*(1-PULSE(0.0,24.0,12.0)) )*rdrink + &
35                (1-rdrink)*mdrinkp(mod(T,24.0))*0.02*DWDOSE*BW)*PULSE(0.0,168,dayon)
36          RAS = KAS*STOM + VAS*STOM/(KMAS+STOM)
37          RSTOM = OWING+  RAOZ-RAS-KSI*STOM   ! Change in stomach (mg/hr)
38          RINT = KSI*STOM -  RFEC - KAI*AINTEST     ! Change in intestines (mg/hr)
39          RLZ = RLIVO*CIZONE  ! Zero-order to liver
40          RAO = RAS + KAI*AINTEST + RLZ    ! Oral absorption (mg/hr)
41          RFEC = KFEC*AINTEST
42          FEC = INTEG(RFEC, 0.0)
43          STOM = INTEG(RSTOM, TODOSE)   ! Amt in stomach (mg)
44          AINTEST = INTEG(RINT, 0.0)  ! Amt in intestines (mg)
45          OralDoseCheck = INTEG(RAO,  0.0)
46
47   ! Arterial Blood
48          RAAB = QC*(CVLU - CAB)
49          AAB = INTEG(RAAB, 0.0)      ! Amount, mg
50          CAB = AAB/VAB             ! Concentration, mg/L
51          AAUCB = INTEG(CAB, 0.0)    ! AUC, hr*mg/L
52
53   ! Fat
54          RF = QF*(CAB - CVF)
55          AF = INTEG(RF, 0.0)  ! Amount, mg
56          CF = AF/VF    ! Concentration, mg/L
57          CVF = CF/PF  ! AUC, hr*mg/L
                                             B-47        DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2   ! Liver
 3          RAL = QL*(CAB - CVL) + RAO - RMETL - RMETL2 - RMETL3
 4          AL = INTEG(RAL, 0.0)  ! Amount, mg
 5          CL = AL/VL          ! Concentration, mg/L
 6          CVL = CL/PL         ! Concentration, mg/L
 7          AUCL = INTEG(CL, 0.0)! ADC, hr*mg/L
 8
 9   ! Liver Metabolism
10          RM ETL = VMAX*CVL/(KM + CVL)
11          METL = INTEG(RMETL, 0.0)
12          RMETL2 = VMAX2*CVL/(KM2 + CVL)
13          METL2 = INTEG(RMETL2, 0.0)
14          RMETL3 = KLL*CVL
15          METL3 = INTEG(RMETL3, 0.0)
16
17   ! Total Amount Metabolized (Formate and Formaldehyde)
18   ! Does not include K1C for human MeOH excretion estimate
19          AMET = METL + METL2 + METL3
20          AMET24 = AMET*24.0/TSTOP
21   ! Total amount metabolized in last tmetf hr of exposure, averaged per day
22          AMETF = INTEG((RMETL+RMETL2+RMETL3)*PULSE(TSTOP-tmetf,TSTOP,tmetf),0.0)* &
23                24.0/tmetf     ! (tmetf = 24.0*metd)
24   ! Chamber concentration  (mg/L)
25          RACh = (Rats*QP*CLEx) - (FRACin*Rats*QP*CCh) - (kLoss*ACh)
26          ACh = INTEG(RACh, AChO)
27
28   ! The following calculation yields an air concentration equal to the
29   ! closed chamber value if a closed chamber run is in place and a
30   ! specified constant air concentration if an open chamber run is in place
31          CCh = ACh*Cizone/VCh
32          CCPPM = CCh*24451/MWMe
33          CLoss = INTEG(kLoss*ACh, 0.0)
34
35   !Lungs
36          RALu = QP*(FRACin*CCh - CLEx) + QC*(CVB - CVLu)
37          ALu = INTEG(RALu, 0.0)
38          CLu = ALu/VLu ! Concentration, mg/L
39          CVLu = CLu/PLu      ! Exiting Concentration, mg/L
40
41   ! Amount Inhaled
42          Rlnh = FRACin*QP*CCh
43          Alnh = INTEG(Rlnh, 0.0)      ! mg per rat
44          AlnhC = Alnh*Rats     ! mg for a group of rats
45
46   ! Amount Exhaled
47          CLEx = CVB/PB       ! Concentration, mg/L
48          RAEx = QP*CLEx
49          AEx = INTEG(RAEx, 0.0)*PULSE(0,TCHNG,TSTOP)    ! Amount, mg per rat
50          AExC = AEx*Rats     ! Amount, mg, for a group of rats
51          AxF = INTEG(RAEx*PULSE(TCHNG,24,24), 0.0)       ! Amount exhaled post-exposure
52
53   ! Rest of Body
54          RAR = QR*(CAB - CVR)
55          AR =  INTEG(RAR, 0.0) ! Amount, mg
56          CR = AR/VR         ! Concentration, mg/L
57          CVR = CR/PR ! Exiting Venous Concentration, mg/L
                                             B-48        DRAFT - DO NOT CITE OR QUOTE

-------
 1         AUCR = INTEG(CR, 0.0)      ! AUC, hr*mg/L
 2
 3   ! Venous Blood (mg)
 4         RURB = K1*CVB*VVB        ! Lumped Clearance from Blood
 5         RAVB = QF*CVF + QL*CVL + QR*CVR + IVR - QC*CVB - RURB
 6         AVB = INTEG(RAVB, 0.0)      ! Amount, mg
 7         CVB = AVB / VVB            ! Concentration, mg/L
 8         AUCB = INTEG(CVB, 0.0)      ! AUC, hr*mg/L (total over entire exposure)
 9         AUCBB = AUCB*24.0/TSTOP   ! Average over exposure, hr*mg/(L*day)
10         AUCBF = INTEG(CVB*PULSE(TSTOP-tmetf,TSTOP,tmetf),0)*24.0/tmetf
11                ! AUCBF = Last tmetf AUC averaged/day (tmetf = 24.0*metd)
12                ! For "steady state" AUC in blood over a day, set exposures to
13                ! several weeks to reach "periodicity", then use AUCBF w/ metd = 7
14
15   ! Bladder compartment, added by PS, U.S. EPA, 10/2008
16         RBL = KBL*ABL       ! Rate of clearance from bladder (mg/hr)
17         ABL = INTEG((RURB-RBL),0.0)! Amount in bladder (mg)
18         RUR= RBL/(BW*0.5e-3)! Urine concentration = rate/[BW*(0.5e-3 L/h/kg BW)]
19         URB =  INTEG(RBL, 0.0) ! Amount cleared to urine, mg
20         URBF = INTEG(RURB*PULSE(TSTOP-tmetf,TSTOP,tmetf),0)*24.0/tmetf
21                ! Amount cleared to urine in last tmetf averaged/day (tmetf = 24.0*metd)
22
23   |*************************  Magg Balance   **************************
24         Tbody = AAB + AF + AL + ALU + AR + AVB + ABL + STOM + AINTEST
25         MetabORCIrd = METB + METL + METL2 + METL3 + AEX + FEC
26         TMass  = Tbody + MetabORCIrd
27         TDose = AinH + INTEG(IVR+DWING+RAOZ+RLZ,0.0) + TODOSE
28         MassBal=100*(TDose-TMass)/(TMass+1e-12)
29         Icompare to TIV, ODOSE, or AINHC
30   ! Check Blood Flows
31         QTOT = QF + QL + QR
32         QRECOV= 100.0*QTOT/QC
33   END  !  End of Derivative
34   TERMT(T.GE.TStop)
35
36   !	Exposure Control	
37   DISCRETE ORALDOSE       ! Stom is amount in stomach
38         IDOSE = DRP(IDS)*DRDOSE*BW
39         STOM = STOM + IDOSE      ! Drinking percent
40         TODOSE = TODOSE + IDOSE
41         IF(IDS.EQ.1)THEN
42                STOM = STOM + ODOSE
43                TODOSE = TODOSE + ODOSE
44         ENDIF
45         IDS = IDS+1
46         IF(IDS.EQ.7)THEN    ! For 6 doses
47                IDS = 1
48                N EWDAY = N EWDAY + 24
49                SCHEDULE ORALDOSE.AT.NEWDAY ! Go to start of the next day
50         ELSE
51                SCHEDULE ORALDOSE.AT.(NEWDAY+DRT(IDS)) ! Go to next drink time
52         ENDIF
53   END  !  OF DISCRETE ORALDOSE
54
55   DISCRETE OFF      ! Turn IN HAL exposure off
56         CIZONE = 0.0
57         IVZONE = 0.0
                                            B-49        DRAFT - DO NOT CITE OR QUOTE

-------
 1          DAY=DAY+1
 2          IF (MULTE) SCHEDULE ON.AT.(DAY*24.0)
 3   END   ! OF DISCRETE OFF
 4
 5   DISCRETE ON
 6          CIZONE=1.0
 7          SCHEDULE OFF.AT.(T+TCHNG)
 8   END   ! OF DISCRETE ON
 9
10   DISCRETE DS1       ! Human at rest
11          ! Equations scheduled for change during simulation repeated here
12          QC = QCCHR*BW**0.75
13          OP = QPCHR*BW**0.75
14          OF = QFC*QC ! QFCHR*QC  ! Equations for alternate flow fractions
15          QL = QLC*QC ! QLCH R*QC  ! But QFC and QLC taken to be 'at rest' values
16          QRC= 1.0-(QFC + QLC)
17          OR = QRC*QC
18   END   ! OF DISCRETE DS1
19
20   DISCRETE DS2       ! Human at work (50VV)
21          ! Equations scheduled for change during simulation repeated here
22          QC = QCCHW*BW**0.75
23          OP = QPCHW*BW**0.75
24          OF = QFC*QC ! QFCHW*QC  ! Equations for alternate flow fractions
25          QL = QLC*QC ! QLCHW*QC  ! But don't seem to work (fit data) well
26          QRC= 1.0-(QFC + QLC)
27          QR = QRC*QC
28   END   ! OF DISCRETE DS2
29
30   END   ! End of Dynamic
31   END   ! End of Program
     B.3.3. acslXtreme procedure (.cmd) file

      File MEOHCBMMfinal.CMD - FOR PBPK MODEL FOR METHANOL
      taken from .cmd file from Ward et al., Edited by KWW - 06/02/96
      Developed for this (CBMM) model - 4/15/15
      Final with Digitized Data - 5/25/05
      Final Version has fast and slow rates of oral absorption
      Version 4 is final version used for simulations
      Final Version 1.10.06
      Beyond this comment, this file is  left "as is" for archival purposes. But most if not all of
      the functions and data sets defined here are replicated and/or replaced in the .m files below.
      Only use these when there is no corresponding .m file.
      - Paul Schlosser, U.S. EPA, Oct.  2008
32
33
34
35
36
37
38
39
40
41
42
43
44   PREPARE T,CVB,MetB
45
46   ! Procedural blocks for general mouse/rat data
47   PROCED CDMICE  ! Anatomic/physiologic data for mice
48   SETBW=0.03, TSTOP=1.5
49   SET IVDose=0, DOSE=0, CONCppm=0
50   SET PL=1.06, PF=0.083, PR=0.66, PB=1350
                                            B-50       DRAFT - DO NOT CITE OR QUOTE

-------
 1   SET QPC=25.4,QCC=25.4,fracin=0.73
 2   SET QLC = 0.25,QFC=0.05
 3   SET KM=12,VmaxC=14.3,KLC=0.0,KAS=2
 4   SET Vmax2c=19,km2=210,KAI=0.22,KSI=l.l
 5   SET VAC = 0.0123,VFC = 0.07,VLC = 0.055
 6   SET VLuC = 0.0073, VVBC = 0.0368
 7   ! Volumes from Brown et al
 8   !Mouse QPC  avg from Brown 29, 24 used in Corley et al and others
 9   ! AVG of measured vent rates by Perkins et al 25.4 L/hr/kgA0.75
10   IBlood volume 4.9% total. As per Brown 25:75 split art:ven
11   IMetab originally from Ward et al - KldC for mice =0
12   END
13
14   PROCED HUMAN
15   SETBW=70
16   SET IVDose=0, DOSE=0, CONCppm=0
17   SET PL=1.06, PR=0.66, fracin=0.75
18   SET VFC=0.214, VLC=0.026,VLUC=0.008
19   SETVAC=0.0198,VVBC=0.0593
20   SET QPC=18.5, QCC=18.5, QLC=0.227, QFC=0.052
21   SET KM=12,VmaxC=l 1,KLC=0.044,KAS=2.0
22   SET KAI=0.22,KSI=1.1
23   SET PB = 1626, PF=0.14
24   SET Vmax2c=0
25   ! Volumes from Brown et al
26   ! QPC from Brown, upper end 13.4 L/hr/kgA0.75
27   INeed higher  for data, 15 L/hr/kgA0.75 used in several published human models
28   IBlood volume 7.9% total. As per Brown 25:75 split art:ven
29   IFrac absorbed from Ernstgard SOT poster + personal communication
30   !Human Partition Coef. equal to mice. Horton et al. used rat
31   lExcept Human Partition Coef blood and fat - from Fiserova-Bergerova and Diaz, 1986
32   ! - but rat values are inconsistent with expected fat partitioning for an alcohol like this
33   ! - for example Pastino and Conolly EtOH model, fat PC =0.1
34   END
35
36   PROCED SDRAT   ! Anatomic/physiologic data for rats
37   SETBW=0.3, TSTOP=1.5
38   SET IVDose=0, DOSE=0, CONCppm=0
39   SETPL=1.6,  PF=0.1,PR=1.3
40   SET KM=45,VmaxC=15,KLC=0.1,KAS=5
41   SET VAC = 0.0185, VFC=0.07, VLC= 0.034,  VLuC=0.005, VVBC=0.0555
42   ! Volumes from Brown et al
43   !PC from horton et al., PF reduced to 0.1 from  Horton's 1.1
44   IBlood volume 7.4% total. As per Brown 25:75 split art:ven
45   IMetab originally from Ward et al - KldC for mice =0
46   I Rat model not calibrated
47   END
48
49   PROCED PREG

                                          B-51       DRAFT - DO NOT CITE OR QUOTE

-------
 1   !For GD 18 mice, BW increased as estimated from Roger's et al
 2   ! Increased VFC as per Corley CRT development review
 3   ! This just to give a WAG as to how data might change from BW and different volume of
 4   distribution
 5   !Not invoked for any PROCs below as the default
 6   ! Liver to 140% of NP
 7   SET BW = 0.055, VFC=0.08,VLC=0.11,VVBC=0.05
 8   END
 9
10   PROCED CLEARIT
11   SET IVDose=0, DOSE=0, CONCppm=0
12   END
13
14   PROCED SHOWIT
15   display Vmaxc,km,klc,pb,pf,pr,pl,kas,fracin
16   END
17
18   [Procedural blocks for all non-pregnant mouse data
19   ! IV
20   PROCED MWARDIV25
21   !Ward etal., TAP 1997
22   ! Figure 2, data from Ward model cmd
23   IData was checked via digitizit - within +1-5% of cmd file
24   CLEARIT
25   CDMICE
26   SET TSTOP=24.0
27   SET IVDOSE=2500., tchng=0.025
28   END
29
30   PROCED PMWARDIV25
31   PLOT /D=MWARDIV25, CVB
32   END
33
34   DATA MWARDIV25(T,CVB)
35   0.08   4481.8
36   0.25   4132.2
37   0.5    3888
38   1.00   3164.8
39   2.0    2303.5
40   4.00   1921.5
41   6      1883.8
42   8      1620
43   12    838
44   18    454.7
45   24    NaN
46   END
47
48   PROCED MWARD95IV25
49   ! Ward etal., FAT 1995

                                          B-52       DRAFT - DO NOT CITE OR QUOTE

-------
0.53
1.06
1.54
3.07
4.07
5.02
6.02
7.02
8.02
9.03
10.03
END
3299.60
3244.54
3190.71
2803.13
2544.36
2237.77
2063.59
1873.10
1521.92
1670.30
1423.12

 1   ! Figure 2
 2   !Data via digitizit
 3   CLEARIT
 4   CDMICE
 5   SET TSTOP=24.0
 6   SET IVDOSE=2500., tchng=0.025
 7   END
 8
 9   PROCED PMWARD95IV25
10   PLOT /D=MWARD95IV25, CVB
11   END
12
13   DATA MWARD95IV25(T,CVB)
14
15
16
17
18
19
20
21
22
23
24
25
26
27   ! Procs for pegnant IV below: MWARDGD9IV25, MWARDGD18IV25, MWARDGD18IV5,
28   MWARDGD18IV1
29   ! Oral
30   PROCED MWARDPO25
31   !Ward etal., FAT 1995
32   ! Figure 2, data from Ward model cmd
33   IData was checked via digitizit - within +1-5% of cmd file
34   CLEARIT
35   CDMICE
36   SET TSTOP=24, DOSE=2500
37   END
38
39   PROCED PMWARDPO25
40   PLOT /D=MWARDPO25, CVB
41   END
42
43   DATA MWARDPO25(T,CVB)
44   0.504 2370
45   0.96  2645
46   1.44  2705
47   1.992 2719
48   2.208 2781
49   3     2704

                                       B-53        DRAFT - DO NOT CITE OR QUOTE

-------
 1   4.008  2370
 2   4.992  2617
 3   6     2516
 4   7.008  2635
 5   7.992  2213
 6   9     2370
 7   10.008 2028
 8   10.992 1916
 9   12    1347
10   13.008 1467
11   13.9921354
12   15    1175
13   16.008 864.3
14   16.992745.2
15   18    422.4
16   19.01  428
17   21    243
18   24    136
19   END
20
21   IProcs for pregnant Oral below: MDORGD8PO15, MWARDGD18PO25
22   ! Inhalation
23   ! QPC set to measured as in Perkins et al., FAT, 1995 for each concentration
24
25   PROCED MPERKIN25
26   !Perkins etal., FAT, 1995
27   !Fig. 2 data in Ward cmd file
28   CLEARIT
29   CDMICE
30   SET TSTOP=24, CONCppm=2500, vchc=5000
31   SET QPC = 29., QCC=29.
32   SET TCHNG=8
33   END
34
35   PROCED PMPERKIN25
36   PLOT /D=MPERKIN25, CVB
37   END
38
39   IThis data from Digitizlt
40   DATA MPERKIN25(T,CVB)
41   2.0    414.0
42   4.0    453.0
43   6.0    586.0
44   8.25  694.0
45   12    282.0
46   16    0.6
47   END
48
49   IThis data from cmd file
                                          B-54       DRAFT - DO NOT CITE OR QUOTE

-------
 1   !DATAMPERKIN25(T,CVB)
 2   11.99        386.49
 3   14.01        617.57
 4   16.00        816.22
 5   18.26        970.27
 6   112.00       393.24
 7   116.0        13.51
 8   END
 9
10   PROCED MPERKIN50
11   !Perkins et al., FAT, 1995
12   !Fig. 2, data in Ward cmd file
13   IData in command file higher than appears in figure
14   CLEARIT
15   CDMICE
16   SET TSTOP=24, CONCppm=5000, vchc=5000
17   SET TCHNG=8, qpc=24.,qcc=24.
18   END
19
20   PROCED PMPERKIN50
21   PLOT /D=MPERKIN5 0, CVB
22   END
23
24   Ithis from Digitizit, Fig 2 Perkins et al
25   DATA MPERKIN50(T,CVB)
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
1
2
O
4
6
8.25
12
16
20
24
END
IThis
644.00
877.00
1340.00
1450.00
2040.00
2290.0
1410.0
583.0
271.0
9.7

data from cmd file
IDATA MPERKIN50(T,CVB)
!1.0
!2.0
!3.0
!4.0
!6.0
!8.3
112.0
116.0
120.0
124.0
906.76
1202.7
1828.38
1986.49
2800
3125.68
1914.86
806.76
367.57
10.81
                                         B-55       DRAFT - DO NOT CITE OR QUOTE

-------
 1   !END
 2
 3   PROCEDMPERKIN100
 4   !Perkins etal., FAT, 1995
 5   !Fig. 2 data in Ward cmd file
 6   INote, Table 6 in Ward paper - max value of 3260 +/- 151
 7   CLEARIT
 8   CDMICE
 9   SET TCHNG=8, CONCppm= 1,0000, tstop=36,vchc=5000
10   SET QPC=21,qcc=21
11   END
12
13   PROCED PMPERKIN100
14   PLOT /D=MPERKIN 100, CVB
15   END
16
17   Ithis from Digitizit, Fig 2 Perkins et al
18   DATAMPERKIN100(T,CVB)
19   2.0    2080.0
20   4.0    2530.0
21   6.0    3350.0
22   8.25  3350.0
23   12    2370.0
24   16    1830.0
25   20    1080.0
26   24    591.0
27   28    44.6
28   END
29
30   !DATAMPERKIN100(T,CVB)
31   ! This from original cmd file
32   !2.0   2809.46
33   !4.0   3405.4
34   !6.0   4528.38
35   !8.3   4524.32
36   112.0  3212.16
37   116.0  2456.76
38   120.0  1439.19
39   124.0  798.65
40   128.0  55.4
41   !END
42
43   ! Procs for Preg mouse Inhalaiton date below: MDOR8IN10,MDOR8IN15
44         !and:MROGGD7IN10, MROGGD6IN1, MROGGD6IN2, MROGGD6IN5,
45   MROGGD6IN10
46
47   ! pregnant mice
48   ! IV
49   PROCED MWARDGD9IV25

                                          B-56        DRAFT - DO NOT CITE OR QUOTE

-------
 1   !Ward etal.,DMD, 1996
 2   !Not used in the manuscript, only in cmd file
 3   CLEARIT
 4   CDMICE
 5   SET TSTOP=24
 6   SET IVDOSE=2500., TINF=0.025
 7   END
 8
 9   PROCED PMWARDGD9IV25
10   PLOT /D=MWARDGD9IV25, CVB
11   END
12
13   DATA MWARDGD9IV25(T,CVB)
14   0.0833 4606.2
15   0.25  4079.5
16   0.5   3489.3
17   1     2939.6
18   2     3447.6
19   4     2605.0
20   6     2690.5
21   8     2574.9
22   12   1506.1
23   18   498.6
24   24.   NaN
25   END
26
27   PROCED PROCED MWARDGD18IV25
28   !Ward etal.,DMD, 1996
29   !Note, Table 6 in Ward paper - max value of 3521+/- 492
30   CLEARIT
31   CDMICE
32   SET TSTOP=24
33   SET IVDOSE=2500., TINF=0.025
34   END
35
36   PROCED PMWARDGD18IV25
37   PLOT /D=MWARDGD18IV25, CVB
38   END
39
40   DATAMWARDGD18IV25(T,CVB)
41   0.0833 4250.0
42   0.25  3445.1
43   0.5   2936.8
44   1.0   2470.5
45   2.0   2528.1
46   4.0   2292.3
47   6.0   2269.4
48   8.0   2057.0
49   12   1805.9
                                        B-57       DRAFT - DO NOT CITE OR QUOTE

-------
 1   18    1482.2
 2   24.0   496.1
 3   END
 4
 5   PROCED MWARDGD18IV5
 6   !Ward etal.,DMD, 1996
 7   INote, Table 6 in Ward paper - max value of 868.8 +/- 53.9
 8   CLEARIT
 9   CDMICE
10   SET TSTOP=6
11   SET IVDOSE=500., TINF=0.025
12   END
13
14   PROCED PMWARDGD181V5
15   PLOT /D=MWARDGD18IV5, CVB
16   END
17
18   DATA MWARDGD 18IV5(T,CVB)
19   0.25   854.7
20   0.5    720.2
21   1.0    624.1
22   2.0    453.2
23   3.0    307.6
24   4.0    217.7
25   4.5    202.6
26   END
27
28   PROCED MWARDGD 18IV1
29   !Ward etal.,DMD, 1996
30   ! Ward Proc GD8, but must be 18 as per PBPK manuscript
31   INote, Table 6 in Ward paper - max value of 252 +/- 12.9
32   liable matches file
33   CLEARIT
34   CDMICE
35   SET TSTOP=4
36   SETIVDOSE=100.
37
3 8   PROCED PMWARDGD 18IV1
39   PLOT /D=MWARDGD18IV1, CVB
40   END
41
42   DATAMWARDGD18IV1(T,CVB)
43   0.25   252
44   0.52   242.2
45   1.0    222.7
46   2      176.4
47   3      134.2
48   3.5    94.41
49   END
                                        B-58       DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2   ! Oral
 3
 4   PROCED MDORGD8PO15
 5   IWard et al., cmd file
 6   INote, Table 6 in Ward paper - max value of 1610 +/- 704
 7   ! Table and file match w/in round off
 8   I Data must be from Dorman
 9   IDorman Teratology, 1995, Fig. 1
10   ! within error for Digitiz data the same
11   CLEARIT
12   CDMICE
13   SETTSTOP=24, DOSE=1500
14   END
15
16   PROCED PMDORGD8PO15
17   PLOT /D=MDORGD8PO15, CVB
18   END
19
20   DATAMDORGD8PO15(T,CVB)
21   1      1609.6
22   2      1331.2
23   4      1241.6
24   8      707.2
25   16    160.0
26   24    38.4
27   END
28
29   PROCED MWARDGD18PO25
30   !Ward etal.,DMD, 1996
31   INote, Table 6 in Ward paper - max value of 3205 +/- 291
32   CLEARIT
33   CDMICE
34   SET TSTOP=24, DOSE=2500
35   END
36
37   PROCED PMWARDGD18PO25
38   PLOT /D=MWARDGD18PO25, CVB
39   END
40
41   I from cmd file, replaced with digitized
42   IDATA MWARDGD 18PO25(T,CVB)
43   10.25  2770.
44   !0.5   3299.
45   !1     3336.
46   12     3502.
47   14     3217.
48   16     2999.
49   !10    2036.
                                        B-59       DRAFT - DO NOT CITE OR QUOTE

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
!12 1832.
!15 949.1
!18 403.5
!21 40.47
!24. 16.03
!END

IDigitizit data
DATA MWARDGD18PO25(T,CVB)
0.5 2024
1 2554
2 3193
4 3002
6 2933
10 1976
12 1922
15 1339
18 1033
21 832
24 580
END

! Inhalation

PROCED MDOR8IN10
! Ward etal., TAP 1997
INote, Table 6 in Ward paper - max value
!Fig 7? Table 6 attributes to Dorman


























of 2080

IDigitizit of Dorman Fig 2 matches cmd file
! actual exposure ppm 9900
CLEARIT
CDMICE



SET TCHNG=6, CONCppm=9900, tstop=36
END

PROCED PMDOR8IN10
PLOT /D=MDOR8IN10, CVB
END

DATA MDOR8IN10(T,CVB)
1 771.2
2 1017.6
4 1788.8
6 2076.8
8 2281.6
16 1152.0
24 268.8
END

















    /-800
B-60       DRAFT - DO NOT CITE OR QUOTE

-------
1
2
4
6
8
16
24
END
1475.2
2486.4
4588.8
7123.2
5888.0
3456.0
1446.4

 1   PROCEDMDOR8IN15
 2   ! Ward etal., TAP 1997
 3   INote, Table 6 in Ward paper - max value of 7136 +/- 736
 4   !Fig 7? Table 6 attributes to Dorman
 5   IDigitizit of Dorman Fig 2 matches cmd file
 6   CLEARIT
 7   CDMICE
 8   SET TCHNG=6, CONCppm=15000, tstop=36
 9   SET vchc=5000000000
10   END
11
12   PROCEDPMDOR8IN15
13   PLOT /D=MDOR8IN15, CVB
14   END
15
16   DATAMDOR8IN15(T,CVB)
17
18
19
20
21
22
23
24
25
26   !Files above provided in cmd file from Ward PBPK model, TAP 1997
27   IFiles below added for this evaluation,
28   ! sources described in proc files and in notebook
29
30   PROCEDMROGGD7IN10
31   ! Rogers et al., Teratology, 1997
32   ! Actual Values kindly Provided by Rogers
33   CLEARIT
34   CDMICE
35   SET TCHNG=7, CONCppm= 1,0000, tstop=36
36   SET vchc=500000000,bw=0.032
37   END
38
39   PROCEDPMROGGD7IN10
40   PLOT /D=MROGGD7IN10, CVB
41   END
42
43   DATA MROGGD7IN10(T,CVB)
44   1      930
45   4      2800
46   6      3360
47   7      3990
48   7.5    3980
49   8      4120
                                         B-61       DRAFT - DO NOT CITE OR QUOTE

-------
 1   9    3270
 2   12    2630
 3   16    1690
 4   26    60
 5   END
 6
 7   PROCED MROGGD6IN1
 8   CLEARIT
 9   IRogers et al., Teratology, 1993
10   IRogers data from GD 6 and 10
11   I In Table 2
12   CDMICE
13   SET TCHNG=7, CONCppm= 1,000, tstop=36,vchc=500000000,bw=0.032
14   END
15
16   PROCED PMROGGD6IN1
17   PLOT /D=MROGGD6IN1, CVB
18   END
19
20   DATA MROGGD6IN1 (T,CVB)
21   7    63
22   7    131
23   END
24
25
26   PROCED MROGGD6IN2
27   I Rogers et al., Teratology, 1993
28   IRogers data from GD 6 and 10
29   I In Table 2
30   CLEARIT
31   CDMICE
32   SET TCHNG=7, CONCppm=2000, tstop=36, vchc=500000000,bw=0.032
33   END
34
35   PROCED PMROGGD6IN2
36   PLOT /D=MROGGD6IN2, CVB
37   END
38
39   DATA MROGGD6IN2(T,CVB)
40   7    487
41   7 641
42   END
43
44   PROCED MROGGD6IN5
45   I Rogers et al., Teratology, 1993
46   IRogers data from GD 6 and 10
47   I In Table 2
48   CLEARIT
49   CDMICE
                                        B-62       DRAFT - DO NOT CITE OR QUOTE

-------
 1   SET TCHNG=7, CONCppm=5000, tstop=36,vchc=500000000,bw=0.032
 2   END
 3
 4   PROCED PMROGGD6IN5
 5   PLOT /D=MROGGD6IN5, CVB
 6   END
 7
 8   DATAMROGGD6IN5(T,CVB)
 9   7     2126
10   7 1593
11   END
12
13   PROCED MROGGD6IN10
14   ! Rogers et al., Teratology, 1993
15   ! Rogers data from GD 6, 10, 15
16   ! In Table 2
17   CLEARIT
18   CDMICE
19   SET TCHNG=7, CONCppm= 1,0000, tstop=36,vchc=500000000,bw=0.032
20   END
21
22   PROCED PMROGGD6IN10
23   PLOT /D=MROGGD6IN10, CVB
24   END
25
26   DATA MROGGD6IN10(T,CVB)
27   7     4653
28   7 4304
29   7 3655
30   END
31
32   IHuman inhalation dta
33
34   PROCED HJOHIN1
35   lErnstgard et al. SOT poster 200 ppm human
36   ! Digitized from Fig 2
37   ! Also personal communication - Ernstgard
38   !QPC from Johanson et al. Scand J. Work Env. 86 =52.6
39   !If Assume value =  alveolar, similar to Astrand '83 value of 56 L/hr/krA0.75
40   IFracin - 50% of total (from poster) -76%
41   !QCC from Corley et al TAP 129, 1994
42   CLEARIT
43   HUMAN
44   SET TCHNG=2, CONCppm= 100, tstop= 16
45   SET QPC=52.6,qcc=26,vchc=500000000
46   END
47
48   PROCED PHJOHIN1
49   PLOT /D=HJOHIN1, CVB
                                         B-63        DRAFT - DO NOT CITE OR QUOTE

-------
 1   END
 2
 3   DATAHJOHIN1(T,CVB)
 4   0.20  0.87
 5   0.46  1.50
 6   0.97  2.31
 7   1.46  3.24
 8   1.91  3.65
 9   2.17  3.52
10   2.50  2.55
11   2.91  2.23
12   3.51  1.59
13   4.01  1.72
14   5.02  0.41
15   6.00  0.50
16   9.24  0.12
17   END
18
19   PROCED HJOHIN2
20   lErnstgard et al. SOT poster 200 ppm human
21   ! Digitized from Fig 2
22   ! Also personal communication - Ernstgard
23   !QPC from Johanson et al. Scand J. Work Env. 86 =52.6
24   ! If Assume value = alveolar, similar to Astrand '83 value of 56 L/hr/krA0.75
25   IFracin - 50% of total (from poster) -75%
26   !QCC from Corley et al TAP 129, 1994
27   CLEARIT
28   HUMAN
29   SET TCHNG=2, CONCppm=200, tstop= 16
30   SET QPC=52.6,qcc=26,vchc=500000000
31   END
32
33   PROCED PHJOHIN2
34   PLOT /D=HJOHIN2, CVB
35   END
36
37   DATA HJOHIN2(T,CVB)
38   0.22  1.63
39   0.49  2.92
40   0.92  4.76
41   1.47  6.30
42   1.90  7.65
43   2.16  6.20
44   2.47  5.49
45   2.91  4.96
46   3.50  3.64
47   4.00  3.43
48   4.99  1.94
49   5.97  1.03
                                           B-64        DRAFT - DO NOT CITE OR QUOTE

-------
 1   8.90   0.21
 2   END
 3
 4   PROCED HOSTERIN2
 5   ! Osterloh et al., JOEM 1996
 6   ! Digitized data provided by EPA
 7   ! Subtracted background from exposure blood levels
 8   CLEARIT
 9   HUMAN
10   SET TCHNG=4, CONCppm=200, tstop= 16
11   SET vchc=500000000, BW=78.2
12   END
13
14   PROCED PHOSTERIN2
15   PLOT/D=HOSTERIN2, CVB
16   END
17
18   DATAHOSTERIN2(T,CVB)
19   0.05   0.54
20   0.25   1.39
21   0.50   1.82
22   0.75   2.28
23   1.00   2.42
24   1.50   2.94
25   2.00   3.37
26   2.50   3.90
27   3.00   4.21
28   3.50   4.61
29   4.00   4.82
30   5.00   2.99
31   6.00   2.30
32   7.00   1.40
33   7.95   1.07
34   END
35
36   PROCED HBATIN82
37   IBatterman et al., Int Arch Occ Health 1998
38   ! Digitized Data
39   CLEARIT
40   HUMAN
41   SET TCHNG=2, CONCppm=800, tstop= 16
42   SET vchc=500000000
43   END
44
45   PROCED PHBATIN82
46   PLOT /D=HBATIN82, CVB
47   END
48
49   DATA HBATIN82(T,CVB)
                                         B-65       DRAFT - DO NOT CITE OR QUOTE

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
2.223 13.658
2.495 13.282
2.742 11.928
3.230 9.456
4.231 6.197
5.247 3.953
6.262 2.325
7.251 1.551
8.216 1.176
END

PROCEDHBATIN81
IBatterman et al., Int Arch Occ Health
! Digitized Data
CLEARIT
HUMAN












1998

SET TCHNG=1, CONCppm=800, tstop=16
SET vchc=500000000
END

PROCEDPHBATIN81
PLOT /D=HBATIN81, CVB
END

DATA HBATIN81(T,CVB)
1.096 6.477
1.398 6.136
1.644 5.345
2.143 4.270
3.178 2.661
4.188 1.307
5.199 0.732
6.266 0.552
7.292 0.356
8.209 0.093
END

PROCED HBATIN830
IBatterman et al., Int Arch Occ Health
! Digitized Data
Ibody weight not provided
CLEARIT
HUMAN


















1998

SET TCHNG=0.5, CONCppm=800, tstop=16
SET vchc=500000000
END

PROCED PHBATIN830
PLOT /D=HBATIN830, CVB




B-66       DRAFT - DO NOT CITE OR QUOTE

-------
 1   END
 2
 3   DATAHBATIN830(T,CVB)
 4   0.579 4.608
 5   0.857 4.685
 6   1.137 4.870
 7   1.650 3.452
 8   2.650 2.082
 9   3.662 0.910
10   4.693 0.316
11   5.713 0.320
12   6.643 0.292
13   7.696 0.547
14   END
15
16   PROCED HSEDIN231
17   ! Sedivec et al., Int Arch Occ Health 1981
18   ! Digitized Data
19   INote, urine volumes not given, these are estimates
20   lurine production of 0.75 mg/hr, this for info purposes only!!!
21   CLEARIT
22   HUMAN
23   SET TCHNG=8, CONCppm=231, tstop=24
24   SET vchc=500000000
25   END
26
27   PROCED PHSEDIN231
28   PLOT /D=HSEDIN231, Metb
29   END
30
31   DATA HSEDIN231 (T,Metb)
32   0.043 0.0042
33   2.174 0.33
34   4.478 0.87
35   6.478 1.46
36   8.522 2.15
37   10.3482.63
38   12.1302.91
39   14.0443.07
40   18.8703.32
41   23.6963.52
42   END
43
44   PROCED HSEDIN157
45   ! Sedivec et al., Int Arch Occ Health 1981
46   ! Digitized Data
47   INote, urine volumes not given, these are estimates
48   lurine production of 0.75 mg/hr, this for info purposes only 11ICLEARIT
49   HUMAN
                                           B-67        DRAFT - DO NOT CITE OR QUOTE

-------
 1   SET TCHNG=8, CONCppm=157, tstop=24
 2   SET vchc=500000000
 3   END
 4
 5   PROCEDPHSEDIN157
 6   PLOT /D=HSEDIN157, Metb
 7   END
 8
 9   DATA HSEDIN157(T,Metb)
10   0.126  0.0038
11   2.204  0.228
12   4.242  0.576
13   6.196  0.975
14   8.326  1.47
15   10.163 1.81
16   12.0942.00
17   14.0162.12
18   18.8966      2.34
19   23.7762.53
20   END
21
22   PROCED HSEDIN78
23   ! Sedivec et al., Int Arch Occ Health 1981
24   ! Digitized Data
25   INote, urine volumes not given, these are estimates
26   lurine production of 0.75 mg/hr, this for info purposes only!!!
27   CLEARIT
28   HUMAN
29   SET TCHNG=8, CONCppm=78, tstop=24
30   SET vchc=500000000
31   END
32
33   PROCED PHSEDIN78
34   PLOT /D=HSEDIN78, Metb
35   END
36
37   DATA HSEDIN78(T,Metb)
38   0.03   0.013
39   2.06   0.189
40   3.96   0.397
41   6.09   0.652
42   8.09   0.820
43   10.11  0.933
44   11.93  1.02
45   13.92  1.09
46   18.89  1.27
47   END
48
49   ! AUC, Cmax estimation procedures
                                          B-68       DRAFT - DO NOT CITE OR QUOTE

-------
 1   Proced mousin
 2   !To determine AUC for 7 hr exposure in mice
 3   CLEARIT
 4   CDMICE
 5   SET TCHNG=7, tstop=24
 6   SET vchc=50000000000
 7   SET CONCppm=l
 8   start /nc
 9   d concppm,AUCB,amet,cvb
10   SET CONCppm=5
11   start /nc
12   d concppm,AUCB,amet,cvb
13   SET CONCppm=10
14   start/nc
15   d concppm,AUCB,amet,cvb
16   SET CONCppm=25
17   start/nc
18   d concppm,AUCB,amet,cvb
19   SET CONCppm=50
20   start /nc
21   d concppm,AUCB,amet,cvb
22   SET CONCppm=75
23   start /nc
24   d concppm,AUCB,amet,cvb
25   SET CONCppm=100
26   start /nc
27   d concppm,AUCB,amet,cvb
28   SET CONCppm=175
29   start /nc
30   d concppm,AUCB,amet,cvb
31   SET CONCppm=208.3
32   start /nc
33   d concppm,AUCB,amet,cvb
34   SET CONCppm=250
35   start/nc
36   d concppm,AUCB,amet,cvb
37   SET CONCppm=325
38   start/nc
39   d concppm,AUCB,amet,cvb
40   SET CONCppm=500
41   start/nc
42   d concppm,AUCB,amet,cvb
43   SET CONCppm=750
44   start /nc
45   d concppm,AUCB,amet,cvb
46   SET CONCppm=l,000
47   start /nc
48   d concppm,AUCB,amet,cvb
49   SET CONCppm=2000
                                         B-69       DRAFT - DO NOT CITE OR QUOTE

-------
 1   start /nc
 2   d concppm,AUCB,amet,cvb
 3   SET CONCppm=2500
 4   start /nc
 5   d concppm,AUCB,amet,cvb
 6   SET CONCppm=5000
 7   start /nc
 8   d concppm,AUCB,amet,cvb
 9   SET CONCppm=l,0000
10   start/nc
11   d concppm,AUCB,amet,cvb
12   SET CONCppm=50000
13   start/nc
14   d concppm,AUCB,amet,cvb
15   END
16
17   Proced mousinC
18   !To determine 7 hr Cmax, note - not at SS
19   CLEARIT
20   CDMICE
21   SET TCHNG=7, tstop=7,VCHC=50000000000
22   SET CONCppm=l
23   start /nc
24   d cone ppm,cvb
25   SET CONCppm=10
26   start /nc
27   d concppm,CVB
28   SET CONCppm=50
29   start /nc
30   d concppm,CVB
31   SET CONCppm=100
32   start /nc
33   d concppm,CVB
34   SET CONCppm=250
35   start/nc
36   d concppm,CVB
37   SET CONCppm=500
38   start/nc
39   d concppm,CVB
40   SET CONCppm=l,000
41   start/nc
42   d concppm,CVB
43   SET CONCppm=2000
44   start /nc
45   d concppm,CVB
46   SET CONCppm=2500
47   start /nc
48   d concppm,CVB
49   SET CONCppm=5000
                                         B-70       DRAFT - DO NOT CITE OR QUOTE

-------
 1   start /nc
 2   d concppm,CVB
 3   SET CONCppm=l,0000
 4   start /nc
 5   d concppm,CVB
 6   SET CONCppm=50000
 7   start /nc
 8   d concppm,CVB
 9   END
10
11   Proced humin
12   ! To determine 24 hr AUC,Cmax at SS for human
13   CLEARIT
14   human
15   SET TCHNG=360, tstop=l,000
16   SET vchc=5000000000, points=48
17   SET Concppm=l
18   Start/nc
19   d concppm,aucBb,cvb
20   SET CONCppm=10
21   start/nc
22   d concppm,aucBb,cvb
23   SET CONCppm=50
24   start /nc
25   d concppm,aucBb,cvb
26   SET CONCppm=100
27   start /nc
28   d concppm,aucBb,cvb
29   SET CONCppm=250
30   start/nc
31   d concppm,aucBb,cvb
32   SET CONCppm=500
33   start/nc
34   d concppm,aucBb,cvb
35   SET CONCppm=625
36   start/nc
37   d concppm,aucBb,cvb
38   SET CONCppm=750
39   start/nc
40   d concppm,aucBb,cvb
41   SET CONCppm=875
42   start /nc
43   d concppm,aucBb,cvb
44   SET CONCppm=l,000
45   start /nc
46   d concppm,aucBb,cvb
47   SET CONCppm=2000
48   start /nc
49   d concppm,aucBb,cvb
                                          B-71        DRAFT - DO NOT CITE OR QUOTE

-------
 1   SET CONCppm=2500
 2   start /nc
 3   d concppm,aucBb,cvb
 4   SET CONCppm=5000
 5   start /nc
 6   d concppm,aucBb,cvb
 7   SET CONCppm=l,0000
 8   start /nc
 9   d concppm,aucBb,cvb
10   SET CONCppm=50000
11   start /nc
12   d concppm,aucBb,cvb
13   END
14
15   Proced humor
16   ! To determine 24 hr AUC
17   ! Oral exposure
18   CLEARIT
19   human
20   SET TCHNG= 1,000, tstop= 1,000
21   SET vchc=5000000000, points=48
22   SETdose=0.1
23   SET ODS=1
24   start /nc
25   d dose,aucBb
26   SET dose=l
27   Start /nc
28   d dose,aucBb
29   SET dose=5
30   start/nc
31   d dose,aucBb
32   SET dose=10
33   start/nc
34   d dose,aucBb
35   SET dose=50
36   start/nc
37   d dose,aucBb
38   SET dose=100
39   start/nc
40   d dose,aucBb
41   SET dose=250
42   start /nc
43   d dose,aucBb
44   SET dose=350
45   start /nc
46   d dose,aucBb
47   SET dose=500
48   start /nc
49   d dose,aucBb
                                           B-72
DRAFT - DO NOT CITE OR QUOTE

-------
 1   SET dose=750
 2   start /nc
 3   d dose,aucBb
 4   SET dose=l,000
 5   start /nc
 6   d dose,aucBb
 7   SET dose=2500
 8   start /nc
 9   d dose,aucBb
10   SET dose=5000
11   start /nc
12   d dose,aucBb
13   S ODS=0
14   END
15
16   !Procedural blocks for all non-pregnant rat data
17   !Not calibrated!!!!!!
18   !Procs from Ward CMD file
19
20   PROCED WARDIV25
21   CLEARIT
22   SDRAT
23   SET TSTOP=48.
24   SET IVDOSE=2500.
25   END
26
27   PROCED PWARDIV25
28   PLOT /D=WARDIV25, CVB
29   END
30
31   DATA WARDIV25(T, CVB)
32   0.072        4849
33   0.168        3926
34   0.24         2965
35   0.504        2836
36   1.008        3248
37   1.992        2589
38   3            2619
39   4.008        2514
40   7.008        2315
41   19.992      1495
42   22.992      1272
43   24          1214
44   25.992      982
45   28.008      957
46   30          860
47   37.992      238
48   39          200
49   40.008      150
                                          B-73       DRAFT - DO NOT CITE OR QUOTE

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
40.992 167
43.008 77
END

PROCED WARDIV1
CLEARIT
SDRAT
SET TSTOP=8
SET IVDOSE=100., tchng=0.016
END

PROCED PWARDIV1
PLOT /D=RGOIV1, CVB
END

DATA WARDIV1 (T,CVB)
0.072 141.7
0.168 121.8
0.24 111.6
0.504 99.7
0.744 97.4
1.008 86.3
1.488 80.3
1.992 58
3 44.4
4.008 22.8
4.992 10.9
6 3.8
7.008 1.4
END

PROCED WARDPO25
CLEARIT
cdmice
SET BW=0.3
SET TSTOP=48
SET DOSE=2500
END

PROCED PWARDPO25
PLOT /D=WARDPO25, CVB
END

DATA WARDPO25(T,CVB)
0.072 862.7
0.168 1243
0.24 1356
0.504 1621
1.008 1641
B-74
DRAFT - DO NOT CITE OR QUOTE

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
1.992 1611
3 1869
4.008 1896
7.008 2181
24 1365
25.992 1081
28.008 921
30 958.4
31.008969.8
45 42.9
46.00827.1
46.992 16.4
48 23.9
49.00841.9
49.992 13.1
52.0082.3
52.992 1
END




















PROCEDWARDPO1
CLEARIT
cdmice
SET BW=0.3



SET DOSE=100, tstop=8
END



PROCED PWARDPO1
PLOT /D=WARDPO1, CVB
END



DATA WARDPO1(T,CVB)
0.072
0.168
0.24
0.504
0.744
1.008
1.488
1.992
O
4.008
4.992
6
7.008
END
85.5
95.6
95.5
91.1
86.6
80.6
71.3
61.1
45.1
27.4
16.4
8.9
4.2

B-75
DRAFT - DO NOT CITE OR QUOTE

-------
     B.3.4. Procedural .m files for reproducing the results in Appendix B and Chapter 3

       B.3.4.1. Key to ACSL Extreme v2.5.0.6.mfiles
 1                               Found in the Runtime Files Folder
 2   CDmice.m     Sets parameters for (CD) mouse simulations
 3   Rogers-mouse-inhal.m      Figure B-2 - Simulations of mouse inhalation exposures from GD 6,
 4                             7, 8 and 10 mice from Rogers et al.,  1993.
 5   PerkinsDorm-mouse-inh.m  Figure B-3 - Simulations of inhalation exposures to MeOH in NP
 6                             mice from Perkins et al. 1995 (8 hr exposures) and GD 8 mice from
 7                             Dorman et al. 1995 (6 hr exposures)
 8   Ward_mouse_GD18.m      Figure B-4 - Oral exposures to MeOH in pregnant and non-pregnant
 9                             mice Data from Dorman et al., 1995 and Ward et al., 1997
10   Ward-mouse-iv.m          Figure B-5 - Simulations of mouse IV exposures to MeOH from
11                             Wardetal., 1997
12   Apaja-mouse-drink.m       Calculates internal doses for mice in Apaja (1980)
13
14   SDrat.m           Sets parameters for Sprague-Dawley (SD) rat simulations
15   F344rat.m          Sets parameters for F344 rat simulations
16   Ward-rat-iv.m      Figure B-10 - Simulations rat IV exposures from Ward et al., 1997 and
17                      Hortonetal., 1992
18   Horton-rat-inhal.m  Figure B-l 1 - Simulations rat inhalation exposures from Horton et al., 1992
19   Ward-rat-oral.m    Figure B-12 - Simulations rat oral exposures from Ward et al., 1997
20   Nedo-rat-inhal-devpmt-rat.m  Figure B-13 - Simulations rat inhalation (bioassay) exposures
21                               (200, 500,  1,000, 2000, & 5000 ppm)
22   Nedo-rat-inhal-cancer.m      Simulations for NEDO F344 rat cancer inhalation study
23   rat-infu-sims.m     Figure B-14 - Simulations rat "oral"  exposures (bioassay doses, but using
24                      liver infusion; for illustration only)
25
26   humanset.m             Sets human MeOH PBPK parameters
27   Sedivec_human_inh.m   Figure B-16 - Simulation of human urinary MeOH elimination
28                          following Inhalation exposures from Sedivec et al. 1981
29   Batterman_human_inh.m   Figure B-17 (upper panel) -  Simulations of human inhalation exposure
30                          data of Batterman et al.  1998
31   Osterloh_human_inh.m  Figure B-17 (lower panel) -  Simulations of human inhalation exposure
32                          data of Osterloh et al. 1996
33   Ernstgard_human_inh.m Figure B-l 8 - Simulations of human inhalation exposures to MeOH
34                          from Ernstgard et al. 2005
35
36   mouse_inh_sim.m   Produces data for Table B-5,  mouse inhalation exposures
37   human_inh_sim.m  Produces data for Table B-5,  human inhalation expsoures
38   human_oral_sim.m  Produces data for Table B-5,  human oral exposures
39   human_drink_compare.m  Figure B-24 and Table B-9 (altenate drinking pattern comparison)
40
41                          Found in the Sensitivity Analysis Files Folder
42   Fig_B-6         Sensitivity of the mouse model to metabolic parameters (e.g., Km and Vmax)
43                    for the inhalation route
44   Fig_B-7         Sensitivity of the mouse model to flow  parameters (e.g., blood flow to liver)
45                    and to the rest-of-body  partition coefficient for the inhalation route

                                              B-76         DRAFT - DO NOT CITE OR QUOTE

-------
 1   Fig_B-8         Sensitivity analysis of the rat model to oral absorption parameters for a bolus
 2                    oral exposure (1,000 mg/kg)

       B.3.4.1. [B.3.4.2.] Code for .m files

 3   % File CDmice.m
 4   % Sets parameters for mouse simulations, MeOH PBPK model
 5   CONCPPM=10; WESITG=0; WEDITG=0; CINT-0.1;
 6   start @nocallback
 7   BW=0.03; TSTOP=24; TCHNG=7; REST=20000; WORK=20000;
 8   IVDOSE=0; DOSE=0; DRDOSE=0; RATS=0; KLOSS=0; LIVRO=0;
 9   PL=1.06; PF=0.083; PR=0.66; PLU=1; PB=1350;
10   QPC=25.4; QCC=25.4; FRACIN=0.73; KFEC=0;
11   QLC=0.25; QFC=0.05;
12   KM=12; VMAXC=14.3; K1C=0.0; KAS=0.0; KLLC=0;
13   VMAX2C=19; KM2=210; KAI=0.5; KSI=5.0;
14   VAC=0.0123; VFC=0.07; VLC=0.055; VLUC=0.0073; VVBC=0.0368;
15   CONCPPM=0.0; IVDOSE=0.0; DOSE=0.0; DWDOSE=0; MULTE=0; RDRINK=1;
16   % Volumes from Brown et al
17   % Mouse QPC avg from Brown 29; 24 used in Corley et al and others
18   % AVG of measured vent rates by Perkins et al 25.4 L/hr/kgA0.75
19   % Blood volume 4.9% total. As per Brown 25:75 split art:ven
20   % Metab originally from Ward et al - KldC for mice = 0
21
22   % use mouselNH_fit-params.m % File contents copied below
23   % Updated parameters as obtained by Paul Schlosser,  U.S. EPA
24   % August 11, 2009 [this file updated]
25
26   % Values generated through parameter estimation script 'mouselNH_fit.m'
27   VMAX2C = 3.222500e+00; KM2 = 660; VMAXC = 19; KM = 5.2; FRACIN = 6.650939e-01;
28
29   % Values generated through parameter estimation script 'mouseor_fit.m'
30   VASC = 1.833246e+03; KSI = 2.2; KAI = 0.33; MASC = 620;
31	
32   % File Rogersjnousejnhal.m (Figure B-2)
33   % Produces MeOH PBPK figures for Rogers' mouse inhalation exposures
34   % Variables in the plot command are case sensitive
35   use CDmice
36          % set mouse parameters
37   %	DATA BLOCKS
38          % These data blocks taken directly from MeOH CBMMv3.cmd
39          % Data for are T (hours), CV (mg/L)
40          % semicolons (";") creates a new line in a data file
41
42   % Rogers et al., Teratology, 1997
43   D7IN10 = [1, 930; 4, 2800; 6, 3360; 7, 3990; 7.5, 3980;
44   8, 4120; 9,  3270; 12, 2630; 16, 1690; 26, 60];
45
46   % Rogers et al.,  Teratology,  1993
47   D6IN1 = [7, 63; 7, 131]; D6IN2 = [7, 487; 7, 641];
48   D6IN5 = [7, 2126; 7, 1593];  D6IN7p5 = [7, 2801; 7, 3455];
49   D6IN10 = [7, 4653; 7, 4304]; D6IN15 = [7, 7720; 7, 7394];
50
51   %	RUN  MODEL
52   RATS=0.0; KLOSS=0.0; % -> open chamber
53   TCHNG=7; CONCPPM=10000; TSTOP=27.0; MULTE=0; BW=0.032;
54   CINT=TSTOP/1000; cs=[]; prepare ©clear T CVB

                                              B-77         DRAFT - DO NOT CITE OR QUOTE

-------
 1    forCONCPPM=[1, 2, 5, 7.5, 10, 15]*1000
 2           start @nocallback
 3           cs=[cs,_cvb];
 4           % Since TSTOP & CINT not changing, assume _t also the same.
 5    end
 6
 7    %	PLOT COM MAN DS
 8           % The rogers.aps file will retain changes made using the plot
 9           % editor as long as the editor is called by clicking the
10           % words  EDIT  PLOT PROPERTIES not the little icon in the
11           % properties dialogue box
12    plot(_t,cs(:,1), _t,cs(:,2), _t,cs(:,3), _t,cs(:,4), _t,cs(:,5), _t,cs(:,6), ...
13           D6IN1(:,1),D6IN1(:,2),D6IN2(:,1),D6IN2(:,2),D6IN5(:,1),D6IN5(:,2), ...
14           D6IN7p5(:,1),D6IN7p5(:,2),D6IN10(:,1),D6IN10(:,2), ...
15           D7IN10(:,1),D7IN10(:,2),D6IN15(:,1),D6IN15(:,2), 'rogers.aps')
16
17    %	WRITE OUT DATA TO A TEXT FILE FOR IMPORTING INTO EXCEL
18       % Cannot save data with different # of rows to the same table.
19    cs=[_t,cs];
20    save cs @file='Rogersplotdata.csv' @format=ASCII @separator=ascii
21	
22    % File: PerkinsDorm-mouse-inh.m (Figure B-3)
23    % Produces MeOH PBPK simulations Perkins 1995 inhalation exposures,
24    % and Ward 1997 (pregnant) and Dorman 1995 for comparison)
25    % Includes all nonpregnant and "early" GD ( open chamber
57    MULTE=0; TSTOP=24; CONCPPM=2500; QPC = 29; QCC=29; TCHNG=8;
                                                B-78         DRAFT - DO NOT CITE OR QUOTE

-------
 1    start @nocallback
 2    Cs25 = _cvb; Ts25 = _t;
 3    CONCPPM=5000; QPC=24; QCC=24; start @nocallback
 4    Cs50 = _cvb; Ts50 = _t;
 5    CONCPPM=10000; TSTOP=36; QPC=21; QCC=21; start @nocallback
 6    Cs100 = _cvb;Ts100 = _t;
 7    use CDmice
 8    RATS=0; KLOSS=0; % -> open chamber
 9    TCHNG=6; CONCPPM=15000; TSTOP=36; start @nocallback
10
11    %	PLOT COM MAN DS
12          % The .aps file will retain changes made using the plot
13          % editor as long as the editor is called by clicking the
14          % words  EDIT PLOT PROPERTIES not the little icon in the
15          % properties dialogue box
16    plot(Ts25, Cs25, Ts50, Cs50, TslOO, Cs100,_t, _cvb, ...
17          Perk25(:,1), Perk25(:,2), Perk50(:,1),  Perk50(:,2), ...
18          Perk100(:,1), Perk100(:,2),Dor815(:,1), Dor815(:,2), 'inhalation.aps')
19
20    %	WRITE OUT DATA TO A TEXT FILE FOR IMPORTING INTO EXCEL
21      % Can't save data with different # of rows to the same table.
22    mytablel = [Ts25, Cs25, Ts50, Cs50, TslOO,  Cs100,_t, _cvb];
23    save mytablel @file='PerkinDormanplotdata.csv' @format=ASCII @separtor=comma
24	
25    %FileWardGD18.m
26    % Creates Figure B-4, including Ward et al 1997 NP and GD 18 mouse data
27    % and Dorman et al 1995 GD 8 mouse data.
28    CDmice
29    TSTOP=25; DOSE=1500; CONCPPM=0; MULTE=0;
30    prepare ©clear T CVB
31    start @nocallback
32    T1=_t;P1=_cvb;
33    DOSE=2500; start @nocallback
34    D15=[1, 1609.6; 2, 1331.2; 4, 1241.6;
35    8, 707.2;  16, 160; 24, 38.4];
36    D25a=[0.5, 2370; 0.96, 2645; 1.44, 2705; 2, 2719;
37    2.2, 2781; 3, 2704; 4, 2370; 5, 2617; 6, 2516;
38    7, 2635; 8, 2213;  9, 2370; 10, 2028; 11, 1916;
39    12, 1347; 13, 1467; 14, 1354; 15, 1175; 16, 864.3;
40    17, 745.2; 18,  422.4; 19, 428; 21, 243; 24, 136];
41    D25b=[0.5, 2024; 1, 2554; 2, 3193; 4, 3002; 6, 2933;
42    10, 1976; 12, 1922; 15, 1339; 18, 1033; 21, 832; 24, 580];
43
44    plot(D15(:,1),D15(:,2),D25a(:,1),D25a(:,2),D25b(:,1),D25b(:,2),...
45          T1,P1,_t,_cvb,"wardgd18plot.aps")
46    % File Ward-mouse-iv.m
47    % M File for reproducing  MeOH PBPK Figure B-5 For WARD iv mouse exposures
48    % (also Ward  Pregnant Includes all nonpregnant and Pregnant)
49	
50    %	DATA BLOCKS
51      %Taken directly from MeOH CBMMv3.cmd, values are [T (hours), CV (mg/L)]
52    % Ward etal, FAT 1995
53    NPIV25=[0.08, 4481.8; 0.25, 4132.2; 0.5, 3888; 1, 3164.8; 2, 2303.5;
54          4, 1921.5; 6, 1883.8; 8, 1620; 12, 838; 18, 454.7; 24, 1.41];
55    %PROCED MWARDGD8IV25
56    GD8IV25=[0.0833, 4606.2; 0.25, 4079.5; 0.5, 3489.3; 1, 2939.6; 2, 3447.6;
57          4, 2605.0; 6, 2690.5; 8, 2574.9; 12, 1506.1; 18, 498.6; 24, 0.554];
                                               B-79         DRAFT - DO NOT CITE OR QUOTE

-------
 1    %!Wardetal., DMD, 1996
 2    GD18IV25=[0.0833, 4250.0; 0.25, 3445.1; 0.5, 2936.8; 1, 2470.5; 2, 2528.1;
 3          4, 2292.3; 6, 2269.4; 8, 2057.0; 12, 1805.9; 18, 1482.2; 24.0, 496.1];
 4    %Wardetal., DMD, 1996
 5    GD18IV5=[0.25, 854.7; 0.5, 720.2; 1, 624.1;
 6          2, 453.2; 3, 307.6; 4, 217.7; 4.5, 202.6];
 7    %Wardetal., DMD, 1996
 8    GD18IV1=[0.25, 252; 0.52, 242.2; 1, 222.7; 2, 176.4; 3, 134.2; 3.5, 94.41];
 9
10    %	RUN MODEL
11    useCDMICE
12    TSTOP=24.0; IVDOSE=2500; TCHNG=0.025; start @nocallback
13    CVs25 = _cvb; Ts25 = _t; TSTOP=6; IVDOSE=500; start @nocallback
14    CVs5 = _cvb; Ts5 = _t; TSTOP=4; IVDOSE=100; start @nocallback
15    CVs1 = _cvb; Ts1 = _t; IVDOSE=200; start @nocallback
16
17    %	PLOT COM MAN DS
18    % The .aps file will retain changes made using the plot
19    % editor as long as the editor is called by clicking the
20    % words EDIT PLOT  PROPERTIES not the little icon in the
21    % properties dialogue box
22    plot(Ts25, CVs25, Ts5, CVs5, Ts1, CVs1, NPIV25(:,1), NPIV25(:,2),...
23          GD8IV25(:,1), GD8IV25(:,2), GD18IV25(:,1), GD18IV25(:,2), ...
24          GD18IV5(:,1), GD18IV5(:,2), GD18IV1(:,1), GD18IV1(:,2), 'iv.aps')
25    plot(_t, _cvb, Ts1, CVs1, GD18IV1(:,1), GD18IV1(:,2), 'ivb.aps')
26
27    %	WRITE OUT DATA TO A TEXT FILE FOR IMPORTING INTO EXCEL
28          % Cant save  data with different # of rows to the same table.
29    mytablel =  [Ts25, CVs25, Ts5, CVs5, _t, _cvb, Ts1, CVs1];
30    save mytablel @file='WardlV.csv' @format=ASCII @separator=comma
31	
32    % File Apaja-mouse-drink.m
33    % Calculates internal doses for mice in Apaja (1980)
34    use CDmice
35    DWDOSE=1; start @nocallback
36    DWDS=[0.045,  550; 0.045, 970; 0.045, 1800;
37          0.040, 560; 0.040, 1000; 0.040. 2100];
38    % Above are BWs and doses for males, then females, from Apaja (1980)
39    ODS=1; TSTOP=24*3*7; MULTE=1; DAYS=6.0; simres=[]; LIVRO=0;
40    PER1=1.5;  DUR1=0.75; PER2=3.0; DUR2=0.5; FNIGHT=0.8; CINT=0.01;
41    prepare @clear T CVB STOM
42    for RDRINK =0 %[1, 0] % 0 -> mouse drinking pattern
43    fori=1:length(DWDS)
44          BW=DWDS(i,1);  DWDOSE=DWDS(i,2); start @nocallback
45          simres=[simres;[TDOSE*(24/TSTOP)/BW,BW,AUCBF,max(_cvb),AMF]];
46    end
47    plot(_t,_cvb)
48    end
49    simres=[simres(:,1)*7/6, simres];
50    simres/100     % Print values to screen (/100)
51    TDOSE*(24/TSTOP)/BW       % Check that final total dose/day is correct
52    save simres @file='Apaja_mouse_drink_sims.csv' @format=ascii @separator=comma
53	
54    % File SDrat.m
55    % Sets parameters for rat simulations, MeOH PBPK model
56    CONCPPM=10; WESITG=0; WEDITG=0; TSTOP=24; TCHNG=6; MULTE=0;
57    REST=20000; WORK=20000;
                                               B-80         DRAFT - DO NOT CITE OR QUOTE

-------
 1    start @nocallback
 2    BW=0.275; TSTOP=24; FRACIN=0.2;
 3    IVDOSE=0; DOSE=0; CONCPPM=0; DRDOSE=0; DWDOSE=0; ODS=0; LIVRO=0;
 4    QCC=16.4; QPC=16.4; QFC=0.07; QLC=0.25;
 5    PL=1.06; PF=0.083; PR=0.66; PB=1350;
 6    KM=6.3; VMAXC=5.0; VMAX2C=8.4; KM2=65;
 7    KLLC=0.0; K1C=0.0;
 8    VAC=0.0185; VFC=0.07; VLC=0.037; VLUC=0.005; VVBC=0.0443;
 9
10    KAS=10.9; KSI=6.8; KAI=0.039; KFEC=0.0; VASC=0;
11    % Just above are linear absorption params fit to 100 mg/kg
12    % oral data, w/ no fecal elimination
13
14    KAS=0.0;  % Below are forsaturable  uptake model
15    % Values  generated through parameter estimation script 'ratoral_fit.m'
16    KSI = 7.4; KAI = 0.051; KFEC = 0.029;  VASC = 5.573125e+03; KMASC = 620;
17	
18    % File F344rat.m: parameters specific to F344 rat
19    % Created by Paul Schlosser, U.S. EPA, Aug. 2009
20    use SDrat
21    VMAXC=0;VMAX2C=22.3; KM2=100;
22	
23    % File: Ward-rat-iv.m
24    % Creates Figure B-9; rat MeOH PBPK model, to simulate
25    % Ward '97 rat iv 2500 & 100 mg/kg  (BW=275, SD)
26    % and Morton '92 iv 100 mg/kg (BW=100, F344)
27    rwi25 =[0.072, 4849; 0.168, 3926; 0.24, 2965; 0.5, 2836; 1, 3248;
28    2, 2589; 3, 2619; 4, 2514; 7, 2315; 20,  1495; 23, 1272; 24, 1214;
29    26, 982; 28, 957; 30, 860; 38, 238; 39, 200;  40,  150; 41, 167; 43, 77];
30
31    % ward  '97 rat iv 100 mg/kg (BW 275, SD)
32    rwi1=[0.072, 141.7; 0.168, 121.8; 0.24,  111.6; 0.5, 99.7; 0.744, 97.4;
33    1, 86.3;  1.488, 80.3; 2, 58; 3, 44.4; 4, 22.8; 5, 10.9; 6, 3.8];
34
35    % horton '92 rat iv 100 mg/kg (BW 220, F344)
36    rhi1=[0.24, 91.13; 0.52, 80.14; 0.90, 70.29; 1.38, 61.22;
37    1.86, 60.63; 2.79, 39.40; 4.30, 26.05; 5.81, 12.51];
38
39    use SDrat
40    prepare @clear T CVB
41    TSTOP=48; IVDOSE=2500; TCHNG=0.016; BW=0.275; CINT=0.1; start @nocallback
42    Twi25 = _t; Cwi25 = _cvb;
43    IVDOSE=100; start @nocallback
44    Twi1 = _t;  Cwi1 = _cvb;
45    BW=0.22; start @nocallback
46
47    plot(Twi25, Cwi25, Twi1, Cwi1, _t, _cvb, rwi25(:,1), rwi25(:,2), ...
48           rwM(:,1), rwi1(:,2), rhi1(:,1), rhi1(:,2), 'rwi2500.aps')
49
50    use F344rat
51    IVDOSE=100; BW=0.22; TCHNG=0.016; CINT=0.1; start @nocallback
52    plot(Twi25, Cwi25, Twi1, Cwi1, _t, _cvb, rwi25(:,1), rwi25(:,2), ...
53           rwi1(:,1), rwi1(:,2), rhi1(:,1), rhi1(:,2), 'rwMOO.aps')
54	
55
56
57    % Creates Figure B-10
                                                B-81        DRAFT - DO NOT CITE OR QUOTE

-------
 1    hi20=[0.46, 18.70; 1, 23.76; 3, 59.73; 6, 80.12
 2    7, 83.25; 9, 53.49; 12, 16.54; 15, 0.91];
 3    hi12=[0.46, 4.89; 1, 8.02; 3, 20.57; 6, 26.63;
 4    7, 16.12; 8, 9.28; 9, 5.23; 10.5, 2.93; 12, 0.98];
 5    hi2=[0.48, 1.2; 3, 3.1; 6, 3.7; 6.47, 2.7;
 6    7, 2.0; 8, 1.6; 9, 1.2];
 7
 8    use F344rat
 9    prepare ©clear T CVB
10    TSTOP=16; CONCPPM=2000; TCHNG=6; BW=0.22; CINT=0.1; start ©nocallback
11    120 = _t; c20 = _cvb;
12    CONCPPM=1200; TSTOP=13; start ©nocallback
13    t12 = _t;c12 = _cvb;
14    CONCPPM=200; TSTOP=10; start ©nocallback
15    t2 = _t; c2 = _cvb;
16    TSTOP=16; CONCPPM=2000; TCHNG=7; start ©nocallback
17
18    plot(t20, c20, t12, c12,12, c2, hi20(:,1), hi20(:,2), ...
19          hi12(:,1), hi12(:,2), hi2(:,1), hi2(:,2), _t, _cvb, 'hi2000.aps')
20	
21    % File: Ward-rat-oral.m
22    % MeOH PBPK model rat simulations for Ward '97 rat oral data
23    % Creates Figre B-11
24    use SDRat
25    KAS=10.9; KSI=6.8; KAI=0.039; KFEC=0.0; VASC=0;
26    BW=0.3; ODS=0; prepare ©clear T CVB
27    DOSE=100; TSTOP=10; start ©nocallback
28    t1=_t;c1=_cvb;
29    DOSE=2500; TSTOP=36; start ©nocallback
30    t2=_t;c2=_cvb;
31
32    KAS=0; use ratoral_fit-params
33    DOSE=100; TSTOP=10; start ©nocallback
34    t1A=_t;c1A=_cvb;
35    DOSE=2500; TSTOP=36; start ©nocallback
36    t2A=_t;c2A=_cvb;
37
38    d1=[0.072, 85.5; 0.168, 95.6; 0.24, 95.5; 0.504, 91.1;
39    0.744, 86.6; 1.008, 80.6; 1.488, 71.3; 1.992, 61.1;
40    3, 45.1; 4.008, 27.4; 4.992,  16.4; 6, 8.9; 7.008, 4.2];
41
42    d25=[0.072, 862.7; 0.168, 1243; 0.24, 1356; 0.504, 1621;
43    1.008, 1641; 1.992, 1611; 3, 1869; 4.008, 1896; 7.008, 2181;
44    24, 1365; 25.992, 1081; 28.008, 921; 30, 958.4; 31.008, 969.8];
45
46    plot(t2,c2, t2A,c2A, d25(:,1),d25(:,2), ...
47          t1,c1, t1A,c1A, d1(:,1),d1(:,2),"wardratoralplot.aps")
48    plot(t2,c2, t2A,c2A, d25(:,1),d25(:,2), ...
49          t1,c1, t1A,c1A, d1(:,1),d1(:,2),"wardratoralplotb.aps")
50	
51    % File: Nedo-rat-inhal-devpmt.m
52    % MeOH PBPK model rat simulations for rat inhalation exposures
53    % 200, 500, 1000, 2000, and 5000 ppm
54    % Internal doses for NEDO developmental inhalation exposures, Sprague-Dawley rats
55    % Creates Figure B-13 ('simres' is tabulated results)
56    use SDRat
57    prepare ©clear T CVB
                                                B-82         DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2    simres=[]; ts=[]; cs=[]; ts2=[]; cs2=[]; TSTOP=24*2*7; CINT=1;
 3    for CONCPPM=[200, 500, 1000, 2000, 5000]
 4          TCHNG=22; MULTE=1; start @nocallback
 5          res=[CONCPPM,max(_cvb),AUCBF,AMF];
 6          ts=[ts,_t]; cs=[cs,_cvb];
 7          TCHNG=TSTOP; MULTE=0; %CONCPPM=22*cp/24;
 8    start @nocallback
 9          simres=[simres;[res,CONCPPM,max(_cvb),AUCBF,AMF]];
10          ts2=[ts2,_t]; cs2=[cs2,_cvb];
11    end
12
13    simres
14    plot(ts(:,2),cs(:,2),ts(:,3),cs(:,3),ts(:,4),cs(:,4), ...
15          ts2(:,2),cs2(:,2),ts2(:,3),cs2(:,3),ts2(:,4),cs2(:,4), ...
16          [2424],[0,95],'fig13.aps')
17    save simres @file='Nedo_devpomt_rat_inhal_sims.csv' @format=ascii @separator=comma
18    cs=[ts(:,1),cs,cs2]; save cs @file='Fig13_sims.csv' @format=ascii @separator=comma
19	
20    % File Nedo-rat-inhal-cancer.m
21    % Simulations for NEDO F344 rat cancer inhalation study
22    use F344rat
23    TCHNG=22.7; TSTOP=5*7*24; MULTE=1;
24    res=[]; CONCPPM=200; prepare ©clear T CVB
25    start @nocallback
26    TCHNG=19.5; ODS=1; cppm=[0,10,100,1000];
27    bwm=[422.1, 418.3, 417.7, 410.0];
28    bwf=[268.7, 270.6, 267.0, 264.9];
29    forsdose=[0,25.4863]
30          fori=1:length(cppm)
31                 CONCPPM=cppm(i);
32                 BW=bwm(i)/1000;
33                 DOSE=sdose/(BWA0.25);
34                 start @nocallback
35                 res=[res;[CONCPPM,TCHNG,BW,AUCBF,max(_cvb),AMF]]
36                 BW=bwf(i)/1000;
37                 DOSE=sdose/(BWA0.25);
38                 start @nocallback
39                 res=[res;[CONCPPM,TCHNG,BW,AUCBF,max(_cvb),AMF]]
40          end
41    end
42
43    save res  @file='Nedo_rat_cancer_sims.csv' @format=ascii @separator=comma
44	
45    % File: rat-infu-sims.m
46    % MeOH PBPK model rat simulations for zero-order liver infusions
47    % Creates Figre B-14
48    use SDRat
49    Iv0=[0.33, 65.9; 0.33, 624.1; 0.34, 2177;
50          0.49, 53.2; 0.50, 524;
51    0.54   1780];
52    % Above are BWs and doses from Soffretti et al. 2002a
53    prepare ©clear T CVB
54    TCHNG=12; MULTE=0; simres=[]; ts=[]; cs=[];
55    fori=1:3
56          BW=lvO(i,1); LIVRO=lvO(i,2);  TSTOP=24; start @nocallback
57          res=[LIVRO,BW,max(_cvb),0,AUCB,0,AMET];
                                               B-83         DRAFT - DO NOT CITE OR QUOTE

-------
 1          TSTOP=84; start @nocallback
 2          res(4)=AUCB; res(6)=AMET; simres=[simres;res];
 3          ts=[ts,_t]; cs=[cs,_cvb];
 4    end
 5    simres/100
 6    plot(ts(:,1),cs(:,1),ts(:,2),cs(:,2),ts(:,3),cs(:,3),'fig14b.aps')
 7    save simres @file='rat_liver-infusion_sims.csv' @format=ascii @separator=comma
 8	
 9    % File: humanset.m
10    % Sets parameters for human simulations. Expects the user to define
11    % metabf = "linear" to use 1st-order metabolism parameters; otherwise
12    % metabf set to "non-linear" and Michaelis-Menten parameters used.
13    BW = 70; FRACIN = 0.8655; IVDOSE=0; DOSE=0; CONCPPM=0; LIVRO=0;
14    PB = 1626; PL = 0.583; %1.06;
15    PF = 0.142; PR = 0.805; %0.66;
16    PLU=1.07;%1.0;
17    VFC = 0.214; VLC = 0.026; VLUC = 0.008;
18    VAC = 0.0198; WBC = 0.0593;
19    QPC = 18.5; QCC = 18.5; QLC = 0.227; QFC = 0.052;
20    KM = 12.76; VMAXC = 0; %VMAXC=11.72; %low KM optimum
21    KM2 = 460; VMAX2C = 0; %VMAX2C=304.5; %high KM optimum
22    KLLC = 60.7; %linear liver metabolism optimum
23    K1C = 0.0397;  KAI = 0.22; KSI = 1.1; KAS = 2.0;
24    RATS=0; KLOSS=0; % constant exposure/no chamber losses
25    QPC=24.0; QCC=16.5; REST=3000; WORK=3000;
26
27    % Below are optimal values for Michaelis-Menten liver metabolism
28    K1C = 0.0342;  KLLC = 0.0; KBL=0.612;
29    KM = 23.7; VMAXC = 33.1;
30
31    % Mouse oral uptake KMASC; others set to match ethanol values
32    % for humans from Sultatos et al. (2004), with VASC set so that
33    % VASC/KMAS = 0.21/h, the Sultatos et al. 1st-order constant,
34    % and KFEC = 0 corresponding to assumed 100% absorption.
35    VASC = 377; KSI = 3.17; KAI = 3.28; KMASC = 620; KFEC=0;
36
37    exist metabf; % check if metabf defined
38    if-ans % If not...
39          metabf = "non-linear"
40    end
41    if metabf=="linear"
42          % Below are optimal values for 1st-order liver metabolism
43          K1C =  0.0373; KLLC = 95.7; KBL=0.564; VMAXC=0.0;
44          % Below are 'no bladder values; uncomment next line to use
45          %K1C = 0.0278; KLLC = 70.3; KBL=1000.0; VMAXC=0.0;
46    else metabf="non-linear";
47    end
48    disp(['Simulation for ',ctot(metabf),' human kinetics']);
49	
50    % File: Sedivec_human_inh.m
51    % Creates MeOH PBPK Figure B-16
52    % For human inhalation exposures, w/ data of Sedivec et al
53
54    o/0	DATA BLOCKS
55          % These data blocks taken directly from MeOH CBMMv3.cmd
56          % Data are T (hours), CV (mg/L), cumulative urinary clearance (mg)
57          % Rounded to 3-4 sig figs
                                                B-84        DRAFT - DO NOT CITE OR QUOTE

-------
 1    % Sedivec etal., Int Arch Occ Health 1981, urine
 2    HS231 = [2, 3.338, 0.1168; 4, 5.776, 0.4358; 6, 7.371, 0.8960;
 3                  8, 8.581, 1.454; 10, 6.576, 1.985; 12, 3.243, 2.328;
 4                  14, 1.32 , 2.488; 19, 0.333, 2.632; 24, 0, 2.661];
 5
 6    HS157 = [2, 2.185, 0.0765; 4, 3.941, 0.291; 6, 4.896, 0.600;
 7                  8, 5.708, 0.971; 10, 4.360, 1.324; 12, 1.738, 1.537;
 8                  14, 0.776, 1.625; 19, 0.231, 1.713; 24, 0, 1.733];
 9
10    HS78 = [2, 0.881, 0.0308; 4, 1.648, 0.1193; 6, 2.285, 0.2570;
11                  8, 2.551, 0.4263; 10, 1.515, 0.5686; 12, 0.708, 0.6464;
12                  14, 0.430, 0.6863; 19, 0.094, 0.7321; 24, 0, 0.7403];
13
14    %	RUN MODEL
15    use humanset
16    TCHNG=8; TSTOP=24; CONCPPM=231;
17    prepare @clear T RU R M ETB
18    start @nocallback
19     ur1 = _metb; t1 = _t; cu1=_rur;, % Save time series for urine MeOHc
20    CONCPPM=157; start @nocallback
21     ur2 = _metb; t2 = _t; cu2= _rur;
22    CONCPPM=78; start @nocallback
23
24    %	PLOT  COM MAN DS
25    plot(t1,ur1,t2,ur2,_t,_metb,HS231(:,1),HS231(:,3),...
26       HS157(:,1),HS157(:,3),HS78(:,1),HS78(:,3), 'sedivic.aps')
27    plot(t1 ,cu1 ,t2,cu2,_t,_rur,HS231 (:,1),HS231 (:,2),...
28       HS157(:,1),HS157(:,2),HS78(:,1),HS78(:,2), 'sedivic2.aps')
29
30    %	WRITE OUT DATA TO A TEXT FILE FOR IMPORTING INTO EXCEL
31       % Cant save data with different # of rows to the same table.
32    mytablel =  [t1 ,ur1 ,cu1 ,t2,ur2,cu2,_t,_metb,_rur];
33    eval(['save mytablel @file=Sedv_fit_KLLC.',num2str(round(KLLC)),'.csv@format=ascii
34    @separator=comma']);
35	
36    % File: Batterman_human_inh.m
37    % Creates MeOH PBPK Figure  B-17 (upper panel)
38    % For human inhalation exposures of Batterman et al 1998
39
40      %These data blocks taken directly from MeOH  CBMMv3.cmd
41      %Data are T (hours), CV (mg/L)
42    % Batterman et al., Int Arch Occ Health 1998
43    HB82=[2, 13.6; 2.25, 13.4; 2.5, 12; 3, 9.6;
44           4, 6.4; 5, 4.1;  6, 2.6; 7, 1.8;  8, 1.4];
45    HB81=[1, 6.5; 1.25, 6.2; 1.5, 5.4; 2, 4.3; 3, 2.8;
46    4, 1.5; 5, 0.94; 6, 0.72; 7, 0.52; 8, 0.23];
47    HB830=[0.5, 4.6; 0.75, 4.7; 1, 4.9; 1.5, 3.5; 2.5, 2.2;
48    3.5, 1; 4.5, 0.52; 5.5, 0.51; 6.5, 0.47; 7.5, 0.68];
49
50    use humanset
51    prepare ©clear T CVB
52    TCHNG=2; CONCPPM=800; TSTOP=16; start @nocallback
53    t2=_t; c2=_cvb; TCHNG=1; start @nocallback
54    t1=_t; c1=_cvb; TCHNG=0.5; start @nocallback
55    t30=_t; c30=_cvb;
56
57    %	PLOT  COMMANDS
                                                 B-85         DRAFT - DO NOT CITE OR QUOTE

-------
 1    plot(t2,c2, t1,d, t30,c30, HB82(:,1),HB82(:,2), ...
 2           HB81(:,1),HB81(:,2),HB830(:,1),HB830(:,2), 'batterman.aps')
 O
 4    %	WRITE OUT DATA TO A TEXT FILE FOR IMPORTING INTO EXCEL
 5       % Cant save data with different # of rows to the same table.
 6    le=1 :min([length(t1),length(t2),length(t30)]);
 7    mytablel = [t2(le),c2(le),t1 (Ie),c1 (Ie),t30(le),c30(le)];
 8    eval(['save mytablel @file=Batter_fit_KLLC.',num2str(round(KLLC)),'.csv '...
 9           '@format=ascii @separator=comma']);
10	
11    % File: Osterloh_human_inh.m
12    % Creates Fig B-17 (lower panel)
13    % Data from Osterloh et al., JOEM 1996
14    % Digitized data provided by EPA (datl)
15    % Subtracted background from exposure blood  levels
16    % by Paul Schlosser, U.S. EPA
17    use humanset
18    BW=78.2;
19    dat1=[0.05, 0.54; 0.25, 1.39; 0.50, 1.82; 0.75, 2.28; 1, 2.42;
20                  1.5, 2.94; 2, 3.37; 2.5, 3.90; 3, 4.21; 3.5, 4.61;
21                  4, 4.82; 5, 2.99; 6, 2.30; 7, 1.40; 8, 1.07];
22    dat=[0.25, 1.183; 0.5, 1.526; 0.75, 1.948; 1, 2.073; 1.5, 2.741;
23                  2, 3.118; 2.5, 3.495; 3, 3.998; 3.5, 4.181; 4, 4.48;
24                  5, 2.790; 6,  1.943; 7, 1.106; 8, 0.687];
25
26    prepare ©clear T CVB
27    TCHNG=4; CONCPPM=200; TSTOP=16; start @nocallback
28    plot(_t,_cvb,dat(:,1),dat(:,2),  'osterloh.aps')
29    mytablel =[_t,_cvb];
30    eval(['save mytablel @file=Oster_fit_KLLC.',num2str(round(KLLC)),'.txt @format=ascii']);
31	
32    % File: Ernstgard_human_inh.m
33    % Creates MeOH PBPK Figure B-18, w/ data of Ernstgard et al 2005a,b
34    % For human inhalation exposures w/ exercise
35    o/0	DATA BLOCKS
36      %These data blocks taken directly from MeOH CBMMv3.cmd
37      %Data are T (hours), CV (mg/L)
38    % Ernstgard et al. SOT poster, 100 ppm & 200 ppm human
39    ernl =[0.20, 0.87; 0.46, 1.50; 0.97, 2.31; 1.46, 3.24;
40           1.91, 3.65; 2.17, 3.52; 2.50, 2.55; 2.91, 2.23; 3.51, 1.59;
41           4.01, 1.72; 5.02, 0.41; 6.00, 0.50; 9.24, 0.12];
42    ern2 =[0.22, 1.63; 0.49, 2.92; 0.92, 4.76; 1.47, 6.30;
43           1.90, 7.65; 2.16, 6.20; 2.47, 5.49; 2.91, 4.96; 3.50, 3.64;
44           4.00, 3.43; 4.99, 1.94; 5.97, 1.03; 8.90, 0.21];
45
46    %	RUN MODEL
47    use humanset
48    QPCHR=QPC; QCCHR=QCC; REST=2.0; WORK=0.0;
49    TCHNG=2.0; CONCPPM=100; TSTOP=10.0; QPCHW=52.6; QCCHW=26.0;
50           %FRACIN=0.9509; %FRACIN=0.9324;
51    prepare T CVB OP QC
52    start @nocallback,
53    cv1 = _cvb; t1  = _t; CONCPPM=200; start @nocallback
54
55    %	PLOT COMMANDS
56    plot(t1,cv1,_t, _cvb,...
57           ern1(:,1),ern1(:,2),ern2(:,1),ern2(:,2), 'ernstgard.aps')
                                                 B-86         DRAFT - DO NOT CITE OR QUOTE

-------
 1    %	WRITE OUT DATA TO A TEXT FILE FOR IMPORTING INTO EXCEL
 2      % Cant save data with different # of rows to the same table.
 3    mytablel = [t1, cv1, _t, _cvb];
 4    eval(['save mytablel @file=Ernst_nofit_KLLC.',num2str(round(KLLC)),'.csv'...
 5          '@format=ascii @separator=comma']);
 6	
 7    % File: mouse_inh_sim.m
 8    % Runs simulations for Table B-5, mouse internal-dose calculations
 9    % from inhalation exposure, over the concentration range specified
10    % in the 'for' statement below.
11    % Results saved to file 'MouselnhalSims.csv'.
12    use CDMice
13    BW=0.03; TCHNG=7; % 7 hr/day exposures
14    TSTOP=240; MULTE=1; % Run for 10 days; multi-day exposure 'on'
15    RATS=0.0;  KLOSS=0.0; % -> open chamber
16    prepare ©clear T CVB
17    CONCPPM=10000; CINT=0.02; start @nocallback
18    % plot(_t,_cvb) % uncomment to see/check that periodicity reached by TSTOP
19    inhres=[]; CINT=0.2;
20    forCONCPPM=[1, 10, 50, 100, 250, 500, 1000, 2000, 5000, 10000]
21          start @nocallback
22          inhres=[inhres;[CONCPPM, AUCBB, max(_cvb), AMET24/(BWA0.75)]];
23    end
24    save inhres @file=MouselnhalSims.csv @format=ASCII @separator=comma
25	
26    % File: human_inh_sim.m
27    % Runs simulations for Table B-5, human internal-dose calculations
28    % from inhalation exposure, over the concentration range specified
29    % in the 'for' statement below.
30    % Results saved to file 'HumanlnhSims_KLLC.#.csv', where # is
31    % value of KLLC used (0 if non-linear/Michaelis-Menten kinetics).
32    % If metab="linear", 1st order kinetics used; otherwise non-linear.
33    use humanset
34    WESITG=0; WEDITG=0; simres=[]; MULTE=0; CINT=1.0; RATS=0.0; KLOSS=0.0;
35    CONCPPM=0; TSTOP=1000; TCHNG=1000; DOSE=0; DWDOSE=0; ODS=1;
36    prepare @clear T CVB STOM
37    start @nocallback
38    forCONCPPM=[1, 10, 50, 100, 250, 500, 1000, 2000, 5000, 10000]
39          start @nocallback
40          simres=[simres;[CONCPPM,AUCBF,max(_cvb),AMF/(BWA0.75)]]
41    end
42    disp(['Simulation for ',ctot(metabf),' human kinetics']);
43    eval(['save simres @file=HumanlnhSims_KLLC.',num2str(round(KLLC)), ...
44                 '.csv @format=ascii @separator=comma']);
45	
46    % File: human_oral_sim.m
47    % Runs simulations for Table B-5, human internal-dose calculations from
48    % oral exposure, over the exposure range specified in the 'for' statement below.
49    % Results saved to file 'Hum_DW_Sims_KLLC.#.csv', where # is value of KLLC
50    % used (0 if non-linear/Michaelis-Menten kinetics).
51    % If metab="linear", 1st order kinetics used; otherwise non-linear.
52    use humanset
53    WESITG=0; WEDITG=0; MULTE=0; CINT=0.1; RATS=0.0; KLOSS=0.0;
54    CONCPPM=0; TSTOP=1000; DWDOSE=0; DOSE=1; ODS=1; DRDOSE=0;
55    prepare @clear T CVB STOM
56    start @nocallback
57    simres=[];
                                               B-87        DRAFT - DO NOT CITE OR QUOTE

-------
 1    forDOSE=[0.1, 1, 10,50, 100,250,500, 1000,2000,5000]
 2          start @nocallback
 3          simres=[simres;[DOSE,AUCBF,max(_cvb)],AMF];
 4    end
 5    disp(['Simulation for ',ctot(metabf),' human kinetics']);
 6    eval(['save simres @file=Hum_DW_Sims_KLLC.',num2str(round(KLLC)),'.csv'...
 7          '@format=ascii @separator=comma']);
 8	
 9    % file human_drink_compare.m
10    % creates Figure B-24 and Table B-9
11    % Created by Paul Schlosser, U.S. EPA, 8/26/09
12    use humanset
13    WESITG=0; WEDITG=0; MULTE=0; CINT=0.1; TSTOP=48; DOSE=0.1; ODS=1; DRDOSE=0;
14    prepare ©clear T CVB
15    start @nocallback
16    T1=_t;C1=_cvb; DOSE=0; LIVRO=0.1; TCHNG=12; MULTE=1; start @nocallback
17    T2=_t; C2=_cvb; LIVRO=0; ODS=0; DOSE=0.1; start @nocallback
18    T3=_t; C3=_cvb; DOSE=0; DRDOSE=0.1; start @nocallback
19    plot(T1 ,C1 ,T2,C2,T3,C3,_t,_cvb,'humoralsim.aps')
20
21    tbl=[];metd = 1.0;
22    fordse=[0.1, 1.0, 10, 100, 250, 500]
23          row=[]; LIVRO=0; DOSE=0; DRDOSE=dse; start @nocallback
24          row=[dse,max(_cvb),AUCBF,AMETF];
25          LIVRO=dse; DOSE=0; DRDOSE=0; start @nocallback
26          row=[row,max(_cvb),AUCBF,AMETF];
27          LIVRO=0; DOSE=dse; DRDOSE=0; start @nocallback
28          tbl=[tbl;[row,max(_cvb),AUCBF,AMETF]];
29    end
30    tbl
31

      B.3.5. Personal Communication from Lena Ernstgard Regarding Human Exposures
      Reported in the Ernstgard and Johanson, 2005 SOT Poster

32    From: Lena Ernstgard [Lena.Ernstgard@imm.ki.se]
33    Sent: Wednesday, March 23, 2005 12:39 AM
34    To: Poet, Torka S
3 5    Subject: RE: Human MeOH poster
36    Hi,
3 7    We measured the ventilation rate and they ought to be similar to those reported by Dr. Johanson at the same
3 8    workload.
39    Sincerly,
40    Lena Ernstgard
41
42    At 18:41 2005-03-22, you wrote:
43
44     Thank you very much. Your net uptake is what we thought. Did you measure ventilation rates?
45
46    Thanks again,
47    Torka
48
49    Torka Poet, PhD
5 0    Center for Biological Monitoring and Modeling
51    Pacific Northwest National Laboratories
52    902 Battelle Blvd.
53    P.O. Box 999, MSIN P7-59

                                               B-88        DRAFT - DO NOT CITE OR QUOTE

-------
 1    Richland, WA 99352
 2    ph: (509)376-7740
 3    fax: (509)376-9449
 4    e-mail: Torka.poet@pnl.gov
 5    (Express Mail Delivery: 790 Sixth Street, Zip Code 99354)
 6
 7           From: Lena Ernstgard [mailto :Lena.Ernstgard(@,imm.ki. sel
 8           Sent: Sunday, March 20, 2005 11:21 PM
 9           To: Poet, Torka S
10           Subject: Re: Human MeOH poster
11           Hi,
12           The manuscript has not been submitted yet, but it will be soon I hope. I will save your mail and send you a
13           copy as soon as possible.
14           When I say % of net uptake, i mean the relative uptake. It is calculated as: cone in exposure chamber -
15           (minus) exhaled cone / (divided by) cone in exposure chamber. I hope you understand how we have done.
16           Sincerly,
17           Lena Ernstgard

      B.3.6.  Personal Communication from Dr. Rogers Regarding Mouse Exposures

18    Jeff Gift, Ph.D.
19    National Center for Environmental Assessment EPA (B243-01) RTF, NC 27711
20    919-541-4828
21    919-541-0245 (fax)
22    gift.jeff@epa.gov
23
24    -— Forwarded by Jeff Gift/RTP/USEPA/US on 04/04/2005 04:31 PM -—
25    John Rogers
26    To:    JeffGift/RTP/USEPA/US@EPA
27    04/04/2005 03:50PM
28    Subject: Re: report(Document link: Jeff Gift)
29
30    Hi Jeff:
31    It's easier just to give you the numbers from that figure in tabular form:
32
33
34    (exposure     Blood [MeOH]   SEM
35    ends @ 7 hr)              (mg/ml)
36
37
38
39
40
41
42
43
44
45
46
47    Whew! Surprised I could find these numbers that fast, I get a little worried when someone asks
48    for 12 year old data, I'm not as organized as I'd like to be,to say the least.
49

                                                B-89        DRAFT - DO NOT CITE OR QUOTE
Time (hr)
(exposure
ends @ 7
1
4
6
7
7.5
8
9
12
16
26


Blood [Me
hr)
0.93
2.80
3.36
3.99
3.98
4.12
3.27
2.63
1.69
0.06

0.05
0.20
0.08
0.13
0.21
0.07
0.16
0.21
0.08
0.02

-------
1
2
3
4
5
6
7
John

John M. Rogers, Ph.D.
Chief, Developmental Biology Branch (MD-67) Reproductive Toxicology Division National
Health and Environmental Effects Research Laboratory Research Triangle Park, NC 27711
T: (919)541-5177
F:(919)541-4017
e-mail: rogers.john@epa.gov

B.3.7. Total MeOH Metabolic Clearance/Metabolites Produced

       Table B-6. Mouse total MeOH metabolic clearance/metabolites produced
       following inhalation exposures3
Exposure concentration
(ppm)
1
10
50
100
250
500
525
550
575
600
625
675
750
875
1,000
2000
5000
1,0000
AUC (mg/L-hr)
1.51E-01
1.53E+00
8.03E+00
1.72E+01
5.38E+01
1.72E+02
1.89E+02
2.09E+02
2.29E+02
2.51E+02
2.74E+02
3.24E+02
4.09E+02
5.77E+02
7.76E+02
5.12E+03
1.73E+04
4.98E+04
Cmax (mg/L)
2.16E-02
2.18E-01
1.15E+00
2.46E+00
7.83E+00
2.64E+01
2.94E+01
3.26E+01
3.62E+01
3.99E+01
4.40E+01
5.30E+01
6.84E+01
9.88E+01
1.34E+02
7.57E+02
2.00E+03
4.60E+03
Total MeOH metabolically cleared
(mg)
1.20E-02
1.20E-01
6.01E-01
1.20E+00
2.99E+00
5.89E+00
6.17E+00
6.45E+00
6.73E+00
7.01E+00
7.28E+00
7.83E+00
8.63E+00
9.93E+00
1.12E+01
2.37E+01
3.77E+01
5.50E+01
    aTotal over a 36-hour period during which mice were exposed for 7 hours to MeOH according to the conditions of
    the dose-response study.
                                              B-90
                                                      DRAFT - DO NOT CITE OR QUOTE

-------
       Table B-7. Human total MeOH metabolic clearance/metabolites produced
       from inhalation exposures3
Exposure concentration (ppm)
1
10
50
100
250
500
625
750
875
1,000
AUC
(mg/L-hr)
0.7142
7.142
35.71
71.42
178.6
357.1
446.4
535.7
625.0
714.2
Cmax (mg/L)
0.0300
0.300
1.498
2.997
7.491
14.98
18.73
22.47
26.22
29.97
Total MeOH metabolically cleared
(mg)
10.23
102.3
511.7
1023
2559
5117
6396
7676
8955
10234
aTotal over a 24-hour period during which humans were exposed continuously to MeOH.
       Table B-8. Human total MeOH metabolic clearance/metabolites produced
       following oral exposures3
Exposure concentration (ms/ks.day)
0.1
1
5
10
50
100
250
AUC (mg/L-hr)
0.3795
3.7954
18.977
37.954
189.8
379.5
948.8
Total MeOH metabolically cleared (mg)
6.2152
62.152
310.8
621.5
3108
6215
15538
aTotal over a 24-hour period during which humans were exposed continuously to MeOH.
Note: MeOH in the model is eliminated via exhalation, metabolism, and urinary excretion (human only). Total
MeOH metabolically cleared approximates total production of down stream metabolites, but as a dose metric is not
equivalent to formaldehyde or formate concentration.
                                           B-91
DRAFT - DO NOT CITE OR QUOTE

-------
     B.3.8. Multiple Daily Oral Dosing for Humans
 1          Current mode simulations of oral exposures to humans use a constant rate of infusion to
 2   the stomach lumen. This approach results in a steady rate of absorption from the stomach equal
 3   to the exposure rate irregardeless of the oral uptake rate constants (assumed equal to the mouse),
 4   hence avoids the difficulty that independent values of these constants are not available for
 5   humans due to a lack of human oral PK data.. A more likely drinking scenario was tested by
 6   using additional code within the model to simulate a 6-times/day drinking schedule, over the
 7   course of 15 h (see code below). The schedule is still an approximation, as it assumes 6 episodes
 8   of drinking, each considered to be  a bolus.  Specifically, it was assumed that humans drank at 0,
 9   3, 5, 8, 11,  and  15 hours from the first ingestion of each day, with the respective fractions of daily
10   consumption being 25, 10, 25, 10,  25, and 5% at those times. The predicted blood
11   concentrations resulting from simulations of six daily boluses, once/day boluses, 12 h/d infusion
12   (zero order), or constant (zero order) are shown in Figure B-24 for a total dose of 0.1 mg/kg.
13   Table B-9 shows PBPK model predicted Cmax, AUC, and Amet (for the last 24 hours of repeated
14   exposures) for humans exposed to MeOH via six daily boluses, 12 h/d infusion, or a single daily
15   gavage.
   0.08-
< 0.07-
•S 0.06 H
 c
.0 0,05-1
4-J
 eo
-b 0,04-1
 cp
 u 0.03-1
 o
 u 0,02^
13
_g 0,01-
CQ
      0-
                                                               Continucrus infusion
                                                               12 h/d infusion
                                                               Single daily bolus
                                                               Six-bolus pattern
                                     10
                              20          30
                               Time (h)
40
            Figure B-24. Simulated human oral exposures to 0.1 mg MeOH/kg/-day
            comparing the first few days for four exposure scenarios: continuous (zero-
            order) infusion; 12 h/d infusion, a single daily bolus, and a pattern of six boluses
            per day (see text).
                                               B-92
                                          DRAFT - DO NOT CITE OR QUOTE

-------
    Table B-9. Repeated daily oral dosing of humans with MeOH"
Dose
(mg/kg)
0.1
1
10
100
250
500
Six daily boluses
^max
(mg/L)
0.0204
0.205
2.16
33.4
182
746
AUC
(mg-h/L)
0.0581
0.584
6.17
109
720
3,290
Amet
(mg)
1.97
19.7
197
1,950
4,400
5,180
12 h/d infusion
^max
(mg/L)
0.0170
0.171
1.83
41.4
237
866
AUC
(mg-h/L)
0.0583
0.586
6.21
125
857
3,540
Amet
(mg)
1.97
19.8
197
1,950
4,420
5,190
Sinsle-dailv bolus
^max
(mg/L)
0.0569
0.579
6.64
95.2
296
903
AUC
(mg-h/L)
0.0584
0.594
7.01
204
1,150
4,250
Amet
(mg)
1.98
19.8
197
1,920
4,440
5,270
*AUC in blood and Amet (amount metabolized) computed from 24-48 h
                            B-93
DRAFT - DO NOT CITE OR QUOTE

-------
                          APPENDIX C.  RfC DERIVATION OPTIONS

      C.I. RFC DERIVATIONS USING THE NEDO METHANOL REPORT (NEDO, 1987)
 1          The BMD approach was utilized in the derivation of potential chronic inhalation
 2    reference values.  In the application of the BMD approach, continuous models in the EPA's
 3    BMDS, version 2.1, were fit to datasets for decreased brain weight in male rats exposed
 4    throughout gestation and the postnatal period to 6 weeks and male rats exposed during gestation
 5    on days 7-17 only (NEDO, 1987). Although there remains uncertainty surrounding the
 6    identification of the proximate teratogen of importance (methanol, formaldehyde, or formate),
 7    the dose metrics chosen for the derivation of RfCs were based on blood methanol levels. This
 8    decision was primarily based on evidence that the toxic moiety is not likely to be the formate
 9    metabolite of methanol (CERHR, 2004), and evidence that levels of the formaldehyde metabolite
10    following methanol maternal and/or neonate exposure would be much lower in the fetus and
11    neonate than in adults.  While recent in vitro evidence indicates that formaldehyde is more
12    embryotoxic than methanol and formate, the high reactivity of formaldehyde would significantly
13    limit its transport from maternal to fetal blood, and the capacity for the metabolism of methanol
14    to formaldehyde is lower in the fetus and neonate versus adults. Further discussions of methanol
15    metabolism, dose metric selection and MOAissues are covered in Sections 3.3, 4.5 and 4.6.

      C.I.I. Decreased Brain Weight in Male Rats Exposed throughout Gestation and into the
      Postnatal Period (NEDO, 1987)
16          The results of NEDO (1987), shown in Table 4-14, indicate that there is not a cumulative
17    effect of ongoing exposure on brain-weight decrements in rats exposed postnatally; i.e., the dose
18    response in terms of percent of control is about the same at 3 weeks postnatal as at 8 weeks
19    postnatal in rats exposed throughout gestation and the FI generation.  However, there does
20    appear to be a greater brain-weight effect in rats exposed postnatally versus rats exposed only
21    during organogenesis (GD7-GD17). In male rats exposed during organogenesis only, there is no
22    statistically significant decrease in brain weight at 8 week after birth at the 1,000 ppm exposure
23    level.  Conversely, in male rats exposed to the same level of methanol throughout gestation and
24    the FI generation, there was an approximately a 5% decrease in brain weights (statistically
25    significant at the/? < 0.01 level).  The extent to which this observation is due to recovery in rats
26    for which exposure was discontinued at birth versus a cumulative effect in rats exposed
27    postnatally is not clear. The fact that male rats exposed to 5,000 ppm methanol only during
28    organogenesis experienced a decrease of brain weight of 10% at 8 weeks postnatal indicates that
29    postnatal exposure is not necessary for the observation of persistent postnatal effects. However,
30    the fact that this decrease was less than the 13% decrease observed in male rats exposed to

                                               C-1            DRAFT - DO NOT CITE OR QUOTE

-------
 1    2,000 ppm methanol throughout gestation, and the 8 week postnatal period indicates that the
 2    absence of postnatal exposure allows for some measure of recovery.
 3           It appears that once methanol exposure is discontinued, continuous biological processes
 4    that are disrupted by exposure, manifesting as decreased brain weight, undergo some recovery
 5    and brain weights begin to return to normal values. This indicates that brain weight is
 6    susceptible to both the level and duration of exposure. Therefore, a dose metric that incorporates
 7    a time component would be the most appropriate metric to use. For these reasons and because it
 8    is more typically used in internal-dose-based  assessments and better reflects total exposure
 9    within a given day, daily AUC (measured for 22-hour exposure/day) was chosen as the most
10    appropriate dose metric for modeling the effects of methanol exposure on brain weights in rats
11    exposed throughout gestation and continuing into the FI generation.
12           Application of the EPA methanol PBPK model (described in Section 3.4) to the NEDO
13    (1987)  study in which developing rats were exposed during  gestation and the postnatal period
14    presents complications that need to be discussed.  The neonatal rats in this study were exposed to
15    methanol gestationally before parturition, as well as lactationally and inhalationally after
16    parturition.  The PBPK model developed by the EPA only estimates internal dose metrics for
17    methanol exposure in NP adult mice and rats. Experimental data indicate that inhalation-route
18    blood methanol kinetics in NP mice and pregnant mice on GD6-GD10 are similar (Dorman
19    et al., 1995; Perkins et al., 1995a,b; Rogers et al.,  1993a,b).  In addition, experimental data
20    indicate that the maternal blood:fetal partition coefficient for mice is approximately 1 (see
21    Section 3.4.1.2).  Assuming that these findings apply for rats, they indicate that pharmacokinetic
22    and blood dose metrics for NP rats are appropriate surrogates for fetal exposure during early
23    gestation. However, as is discussed to a greater extent in Section  5.3, the additional routes of
24    exposure presented to the pups in this study (lactation and inhalation) present uncertainties that
25    make it reasonable to assume that average blood levels in pups in the NEDO report are also
26    greater than those of the dam. However, it is  also reasonable to assume that any differences seen
27    between the pups and dams would also be seen in mothers and human offspring.  Therefore, the
28    presumed differences between pup and dam blood methanol levels are deemed relatively
29    inconsequential, and the PBPK model-estimated adult blood methanol levels are assumed to be
30    appropriate dose metrics for the purpose of this analysis.
31           The first step in the current analysis is to convert the inhalation doses, given as ppm
32    values from the studies, to an internal dose surrogate  or dose metric using the EPA PBPK model
33    (see Section 3.4). Predicted AUC values for methanol in the blood of rats and humans are
34    summarized in Table C-l.
35
                                                C-2            DRAFT - DO NOT CITE OR QUOTE

-------
          Table C-l. EPA's PBPK model estimates of methanol blood levels (AUC) in rats
          following inhalation exposures
Exposure level (ppm)
500
1,000
2000
Methanol in blood AUC (hr x mg/L)a
in rats
79.2
226.7
967.8
      aAUC values were obtained by simulating 22 hour/day exposures for 5 days and calculated for the last 24 hours of
      that period.
 1          The current BMD technical guidance (U.S. EPA, 2000b) suggests that in the absence of
 2    knowledge as to what level of response to consider adverse, a change in the mean equal to
 3    1 control S.D. from the control mean can be used as a BMR for continuous endpoints.  However,
 4    it has been suggested that other BMRs, such as 5% change relative to estimated control mean,
 5    are also appropriate when performing BMD analyses on fetal weight change as a developmental
 6    endpoint (Kavlock et al., 1995).  Therefore, in this assessment, both a 1 control mean S.D.
 7    change and a 5% change relative to estimated control mean were considered.  All models were fit
 8    using restrictions and option settings suggested in the EPA BMD technical guidance document
 9    (U.S. EPA, 2000b).

      C.I.2. BMD Approach with a BMR of 1 Control Mean S.D. -Gestation and into the
      Postnatal Period (NEDO, 1987)
10          A summary of the results most relevant to the development of a POD using the BMD
11    approach (BMD, BMDL, and model  fit statistics) for decreased brain weight at 6 weeks in male
12    rats exposed to methanol throughout gestation and continuing into the FI generation, with a
13    BMR of 1 control mean S.D, is provided in Table C-2. The 6 week male brain weight responses
14    were chosen because they resulted in lower BMD and BMDL estimates than male responses at 3
15    and 8 weeks and female responses at any time point (data not shown). Model fit and was
16    determined by statistics (AIC and $ residuals of individual dose groups) and visual inspection,
17    as recommended by EPA (2000b). There is a 2.5-fold range of BMDL estimates from adequately
18    fitting models, indicating considerable model dependence. In addition, the fit of the Hill and
19    more complex Exponential models is better than the other models  in the dose region of interest
20    as indicated by a lower scaled residual at the dose group closest to the BMD (0.09 versus -0.67
21    or -0.77) and visual inspection.  In accordance with EPA BMD Technical Guidance (EPA,
22    2000b), the BMDL from the Hill model  (bolded), is selected  as the most approriate basis for an
23    RfC derivation because it results in the lowest BMDL from among a broad range of BMDLs and
24    provides a superior fit in the low dose region nearest the BMD. The BMDLiso was determined
25    to be 90.9 hr x mg/L using the 95% lower confidence  limit of the dose-response curve  expressed
26    in terms of the AUC for methanol in blood.
                                               C-3            DRAFT - DO NOT CITE OR QUOTE

-------
     Table C-2. Comparison of BMDiSD results for decreased brain weight in male rats at 6
     weeks of age using modeled AUC of methanol as a dose metric
Model
Linear
2nd degree polynomial
3rd degree polynomial
Power
Hillb
Exponential 2
Exponential 3
Exponential 4
Exponential 5
BMD1SD (AUC,
hr x mg/L) a
278.30
278.30
278.30
278.30
170.57
260.94
260.94
172.08
172.08
BMDL1SD
(AUC,
hr x mg/L)a
225.30
225.30
225.30
225.30
90.93
209.10
209.10
96.93
96.93
^-value
0.5376
0.5376
0.5376
0.5376
0.8366
0.612
0.612
0.8205
0.8205
AICC
-203.84
-203.84
-203.84
-203.84
-203.04
-204.10
-204.10
-203.10
-203.10
Scaled residual"1
-0.77
-0.77
-0.77
-0.77
0.09
-0.67
-0.67
0.09
0.09
aThe BMDL is the 95% lower confidence limit on the AUC estimated to decrease brain weight by 1 control mean
S.D. using BMDS model options and restrictions suggested by EPA BMD technical guidance (U.S. EPA, 2000b).
bThere is a 2.5-fold range of BMDL estimates from adequately fitting models, indicating considerable model
dependence. In addition, the fit of the Hill and more complex Exponential models is better in the dose region of
interest as indicated by a lower scaled residual at the dose group closest to the BMD (0.09 versus -0.67 or -0.77)
and visual inspection. Thus, in accordance with EPA BMD Technical Guidance (EPA, 2000), the BMDL from the
Hill model (bolded) is considered the most appropriate POD for us in an RfC derivation.
°AIC = Akaike Information Criterion = -2L + 2P, where L is the log-likelihood at the maximum likelihood
estimates for the parameters, and P is the number of modeled degrees of freedom (usually the number of parameters
estimated).
d%2d residual (measure of how model-predicted responses deviate from the actual data) for the dose group closest to
the BMD scaled by an estimate of its S.D.  Provides a comparative measure of model fit near the BMD.  Residuals
that exceed 2.0 in absolute value should cause one to question model fit in this region.

Source: NEDO(1987).
 1
 2
 O
 4
 5
 6
 7
 8
 9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
         Hill Model.  (Version: 2.14;   Date:  06/26/2008)
         Input Data File:  C:\Usepa\BMDS21\Data\Methanol\NEDO\Gest-Fl\hilm-6wk-brwHil-
Restrict.(d)
         Gnuplot Plotting File:   C:\Usepa\BMDS21\Data\Methanol\NEDO\Gest-Fl\hilm-6wk-
brwHil-Restrict.pit
                                                   Mon Aug 24  13:08:09 2009
 BMDS Model Run



   The  form of the  response function is:


   Y[dose]  = intercept + v*dose/xn/(k^n +  dose^n)
   Dependent variable = Mean
   Independent variable = Dose
   rho  is set to  0
   v  is set to -0.5
   Power parameter  restricted to be greater than 1
   A  constant variance model is  fit


   Total number of  dose groups = 4
   Total number of  records with  missing values = 0
   Maximum number of iterations  = 250
   Relative Function Convergence has been  set to: le-008
                                             C-4
                                                                   DRAFT - DO NOT CITE OR QUOTE

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
T^ /
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
Parameter Convergence has been set to: le-008



Default Initial Parameter Values
alpha = 0.00539333
rho = 0 Specified
intercept = 1.78
v = -0.5 Specified
n = 1.2699
k = 924.206


Asymptotic Correlation Matrix of Parameter Estimates

( *** The model parameter (s) -rho -v
have been estimated at a boundary point, or have
the user,
and do not appear in the correlation matrix )

alpha intercept n k

alpha 1 1.7e-008 -1.7e-008 -2.9e-009

intercept 1.7e-008 1 -0.56 -0.73

n -1.7e-008 -0.56 1 0.15

k -2.9e-009 -0.73 0.15 1



Parameter Estimates

95.0% Wald
Variable Estimate Std. Err. Lower Conf. Limit
alpha 0.00496114 0.0010023 0.00299667
intercept 1.77806 0.0193152 1.7402
n 1.07212 0.217805 0.645228
k 915.874 179.818 563.437



Table of Data and Estimated Values of Interest

Dose N Obs Mean Est Mean Obs Std Dev Est Std Dev


0 12 1.78 1.78 0.07 0.0704
79.2 12 1.74 1.74 0.09 0.0704
226.7 11 1.69 1.69 0.06 0.0704
967.8 14 1.52 1.52 0.07 0.0704



Model Descriptions for likelihoods calculated


Model Al: Yij = Mu(i) + e(ij)
Var{e(ij)} = Sigma/x2

Model A2 : Yij = Mu(i) + e(ij)
Var{e(ij)} = Sigma (i) ^2

Model A3: Yij = Mu(i) + e(ij)
Var{e(ij)} = Sigma/x2
Model A3 uses any fixed variance parameters that
















been specified by

















Confidence Interval
Upper Conf. Limit
0.00692562
1.81591
1.49901
1268.31





Scaled Res.


0.0956
-0.21
0.159
-0.0355















C-5
DRAFT - DO NOT CITE OR QUOTE

-------
1
2
3
4
5
6
7
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
were specified by the user

Model R: Yi = Mu + e(i)
Var{e (i) } = Signal








Likelihoods of Interest
Model Log (likelihood) #
Al 105.539862
A2 106.570724
A3 105.539862
fitted 105.499930
R 77.428662


Explanation of Tests

Param' s AIC
5 -201.079724
8 -197.141449
5 -201.079724
4 -202.999859
2 -150.857324




Test 1: Do responses and/or variances differ among Dose levels?
(A2 vs. R)

Test 2: Are Variances Homogeneous? (Al vs A2)
Test 3: Are variances adeguately modeled? (A2 vs. A3)
Test 4: Does the Model for the Mean Fit?
(Note: When rho=0 the results of Test 3

Tests of Interest

Test -2*log (Likelihood Ratio) Test

Test 1 58.2841 6
Test 2 2.06173 3
Test 3 2.06173 3
Test 4 0.0798644 1

The p-value for Test 1 is less than .05.
(A3 vs. fitted)
and Test 2 will be the same.)



df p-value

<.0001
0.5597
0.5597
0.7775

There appears to be a
difference between response and/or variances among the dose levels
It seems appropriate to model the data

The p-value for Test 2 is greater than .1.
model appears to be appropriate here


The p-value for Test 3 is greater than .1.
to be appropriate here

The p-value for Test 4 is greater than .1.
to adeguately describe the data


Benchmark Dose Computation

Specified effect = 1

Risk Type = Estimated standard

Confidence level = 0.95

BMD = 169.597



A homogeneous variance



The modeled variance appears


The model chosen seems







deviations from the control mean





C-6
DRAFT - DO NOT CITE OR QUOTE

-------
                BMDL =
                             103.912
                                     Hill Model with 0.95 Confidence Level
      o
      Q.
     o:
     c
     (0
     OJ
              1.85
               1.8
              1.75
               1.7
              1.65
               1.6
              1.55
               1.5
              1.45
                          Hill
                       BMDL     BMD
                                 200
                                400          600
                                    dose
800
1000
       09:05 08/24 2009
         Figure C-l. Hill model, BMR of 1 Control Mean S.D. - Decreased Brain weight in
         male rats at 6 weeks age versus AUC, Fl Generation inhalational study
           Source: NEDO(1987).

1          Once the BMDLiso was obtained in units of hr x mg/L, it was used to derive a chronic
2   inhalation reference value. The first step is to calculate the HEC using the PBPK model
3   described in Appendix B. An algebraic equation is provided (Equation  1 of Appendix B) that
4   describes the relationship between predicted methanol AUC and the human equivalent inhalation
5   exposure concentration (HEC) in ppm.
6
7
BMDLHEC (ppm)= 0.0224*BMDLisD+(1334*BMDLisD)
  BMDLHEC (ppm)= 0.0224*374.7+(1334*374.7)/(794+ 90.9) = 139 ppm
             Next, because RfCs are typically expressed in units of mg/m , the HEC value in ppm
    was converted using the conversion factor specific to methanol of 1 ppm =1.31 mg/m3:
                                             C-7
                                              DRAFT - DO NOT CITE OR QUOTE

-------
                            HEC (mg/m3) = 1.31 x 139 ppm = 182 mg/m3
 2             Finally, this HEC value was divided by a composite 100-fold UF (3 for uncertainty
 3    associated with animal to human differences, 10 for consideration of human variability, and 3 for
 4    database deficiencies) to obtain the chronic inhalation reference value:

 5                          RfC (mg/m3) = 182 mg/m3 -100 = 1.8 mg/m3

      C.I.3. BMD Approach with a BMR of 0.05 Change Relative to Estimated Control Mean -
      Gestation and into the Postnatal Period (NEDO, 1987)
 6          A summary of the results most relevant to the development of a POD using the BMD
 7    approach (BMD, BMDL, and model fit statistics) for decreased brain weight at 6 weeks in male
 8    rats exposed to methanol throughout gestation and continuing into the FI generation, with a
 9    BMR of 0.05 change relative to estimated control mean, is provided in Table C-3. The 6 week
10    male  brain weight responses were chosen because they resulted in lower BMD and BMDL
11    estimates than male responses at 3 and 8 weeks and female responses at any time point (data not
12    shown). Model fit was determined by statistics (AIC and $ residuals of individual dose groups)
13    and visual inspection, as recommended by U.S. EPA (2000b). There is a 2.4-fold range of
14    BMDL estimates from adequately fitting models, indicating considerable model dependence. In
15    addition, the fit of the Hill and more complex Exponential models is better than the other models
16    in the dose region of interest as indicated by a lower scaled residual  at the dose group closest to
17    the BMD (0.09 versus -0.67 or -0.77) and visual inspection. In accordance with EPA BMD
18    Technical Guidance (EPA, 2000b), the BMDL from the Hill model (bolded), is selected as the
19    most  approriate basis for an RfC derivation because it results in the lowest BMDL from among a
20    broad range of BMDLs and provides a superior fit in the low dose region nearest the BMD.  The
21    BMDLos was determined to be 123.9 hr x mg/L, using the 95% lower confidence limit of the
22    dose-response curve expressed in terms of the AUC for methanol in  blood.
                                               C-8           DRAFT - DO NOT CITE OR QUOTE

-------
     Table C-3. Comparison of BMD0s results for decreased brain weight in male rats at 6
     weeks of age using modeled AUC of methanol as a dose metric
Model
Linearb
2nd degree polynomial
3rd degree polynomial
Power
Hill
BMD05 (AUC,
hr x mg/L)a
344.49
344.49
344.49
344.49
223.18
BMDLos (AUC,
hr x mg/L)a
297.92
297.92
297.92
297.92
123.87
/7-value
0.5376
0.5376
0.5376
0.5376
0.8366
AICC
-203.84
-203.84
-203.84
-203.84
-203.04
Scaled Residual"1
-0.77
-0.77
-0.77
-0.77
-0.09
Exponential! 325.82 278.26 0.612 -204.10 -0.67
Exponentials 325.82 278.26 0.612 -204.10 -0.67
Exponential 4 223.94 129.97 0.8205 -203.10 0.09
Exponentials 223.94 129.97 0.8205 -203.10 0.09
The BMDL is the 95% lower confidence limit on the AUC estimated to decrease brain weight by 5% using BMDS
model options and restrictions suggested by EPA BMD technical guidance (U.S. EPA, 2000b).
bThere is a 2.4-fold range of BMDL estimates from adequately fitting models, indicating considerable model
dependence.  In addition, the fit of the Hill and more complex Exponential models is better in the dose region of
interest as indicated by a lower scaled residual at the dose group closest to the BMD (0.09 versus -0.67 or -0.77)
and visual inspection. Thus, in accordance with EPA BMD Technical Guidance (EPA, 2000), the BMDL from the
Hill model (bolded) is considered the most appropriate POD for us in an RfC derivation.
°AIC = Akaike Information Criterion = -2L + 2P, where L is the log-likelihood at the maximum likelihood
estimates for the parameters, and P is the number of modeled degrees of freedom (usually the number of parameters
estimated).
dj^d residual (measure of how model-predicted responses deviate from the actual data) for the dose group closest to
the BMD scaled by an estimate of its S.D. Provides a comparative measure of model fit near the BMD. Residuals
that exceed 2.0 in absolute value should cause one to question model fit in this region.

Source: NEDO(1987).
1
2
O
4
5
6
7
c
O
9
10
1 1
1 1
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
Hill Model. (Version: 2.14; Date: 06/26/2008)

Input Data File: C:\Usepa\BMDS21\Data\Methanol\NEDO\Gest-Fl\hilm-6wk-brwHil-
Restrict. (d)

Gnuplot Plotting File: C:\Usepa\BMDS21\Data\Methanol\NEDO\Gest-Fl\hilm-6wk-
brwHil-Restrict .pit
Mon Aug 24


BMDS Model Run


The form of the response function is:

Y[dose] = intercept + v*dose/xn/ (k^n + dose^n)


Dependent variable = Mean
Independent variable = Dose
rho is set to 0
Power parameter restricted to be greater than 1
A constant variance model is fit

Total number of dose groups = 4
Total number of records with missing values = 0
Maximum number of iterations = 250
Relative Function Convergence has been set to: le-008
Parameter Convergence has been set to: le-008


09:40:25 2009






















                                               C-9
DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
                  Default  Initial  Parameter Values
                          alpha
                            rho
                      intercept
                             v
                             n
                             k
                         0.00539333
                                  0
                               1.78
                              -0.26
                            1.08282
                            401.076
Specified
the user,
Asymptotic Correlation Matrix of  Parameter Estimates


 ( *** The model parameter(s)   -rho     -n
      have been estimated at  a boundary point, or have been specified by


      and do not appear in the correlation matrix )


                                                 k


                                          6.2e-009


                                             -0.64


                                             -0.99


                                                 1

alpha
intercept
V
k
alpha
1
4e-009
-1.2e-008
6.2e-009
intercept
4e-009
1
0.54
-0.64
V
-1.2e-008
0.54
1
-0.99
                         Parameter  Estimates
                                                95.0% Wald Confidence Interval
Variable
alpha
intercept
        v
        n
        k
              Estimate
              0.00495736
              1.77823
             -0.600699
              1
              1284.65
                   Std.  Err.
                    0.00100154
                  0.0184949
                    0.340055
                   NA
                     1317.9
Lower Conf.  Limit
       0.00299438
       1.74198
        -1.2672
                                                 -1298.39
               Upper  Conf.  Limit
                      0.00692034
                      1.81448
                     0.0657974


                      3867.68
NA - Indicates that this  parameter has hit a bound
     implied by some inequality  constraint and thus
     has no standard error.
     Table of Data and Estimated Values of Interest


 Dose       N    Obs Mean     Est Mean   Obs Std Dev  Est Std Dev   Scaled Res.
0
79.2
226.7
967.8
12
12
11
14
1.78
1.74
1.69
1.52
1.78
1.74
1. 69
1.52
0.07
0.09
0.06
0.07
0.0704
0.0704
0.0704
0.0704
0.0872
-0.165
0.0884
-0.00677
 Model Descriptions for likelihoods  calculated
 Model Al:         Yij  = Mu(i)  +  e(ij)
           Var{e(ij)}  = Sigma/x2


 Model A2:         Yij  = Mu(i)  +  e(ij)
           Var{e(ij)}  = Sigma(i)^2
                                              C-10
                                                        DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
 Model A3:        Yij  = Mu(i)  + e(ij)
           Var{e(ij)}  = Signal
     Model A3 uses any fixed variance  parameters  that
     were specified by the user


 Model  R:         Yi  = Mu + e(i)
            Var{e (i) }  = Signal
                       Likelihoods  of  Interest
            Model      Log(likelihood)
             Al          105.539862
             A2          106.570724
             A3          105.539862
         fitted          105.518584
              R           77.428662
                   Explanation of Tests


 Test 1:  Do responses and/or variances  differ  among Dose levels?
          (A2 vs.  R)
 Test 2:  Are Variances Homogeneous?  (Al vs A2)
 Test 3:  Are variances adeguately modeled?  (A2 vs. A3)
 Test 4:  Does the Model for the Mean Fit?  (A3  vs.  fitted)
 (Note:   When rho=0 the results of Test  3 and Test  2 will be the same.)
# Param' s
5
8
5
4
2
AIC
-201.079724
-197.141449
-201.079724
-203.037168
-150.857324
   Test


   Test 1
   Test 2
   Test 3
   Test 4
                     Tests of Interest
-2*log(Likelihood Ratio)   Test  df
            58.2841
            2.06173
            2.06173
          0.0425562
 p-value


<.0001
0.5597
0.5597
0.8366
The p-value for Test 1 is less than .05.   There  appears to be a
difference between response and/or variances  among  the dose levels
It seems appropriate to model  the  data


The p-value for Test 2 is greater  than .1.  A homogeneous variance
model appears to be appropriate here
The p-value for Test 3 is greater than .1.   The modeled variance appears
 to be appropriate here


The p-value for Test 4 is greater than .1.   The model chosen seems
to adeguately describe the data
        Benchmark Dose Computation


Specified effect =          0.05


Risk Type        =     Relative risk


Confidence level =           0.95


             BMD =        223.178
                                              C-ll
                                                        DRAFT - DO NOT CITE OR QUOTE

-------
                BMDL =
                             123.872
                                     Hill Model with 0.95 Confidence Level
      o
      Q.
     o:
     c
     (0
     OJ
              1.85
               1.8
              1.75
               1.7
              1.65
               1.6
              1.55
               1.5
              1.45
                          Hill
                         BMDL
BMD
                                 200
          400          600
              dose
800
1000
       09:40 08/24 2009
         Figure C-2. Hill model, BMR of 0.05 relative risk - Decreased Brain weight in male
         rats at 6 weeks age versus AUC, FI Generation inhalational study.
           Source:  NEDO(1987).

1          Once the BMDLos was obtained in units of hr  x mg/L, it was used to derive a chronic
2   inhalation reference value.  The first step is to calculate the HEC  using the PBPK model
3   described in Appendix B. An algebraic equation is provided (Equation 1 of Appendix B) that
4   describes the relationship between predicted methanol AUC and the human equivalent inhalation
5   exposure concentration (HEC) in ppm.

6              BMDLHEC (ppm)= 0.0224*BMDL05+(1334*BMDLo5)/(794+ BMDL05)
7             BMDLHEC (ppm)= 0.0224*503.0 + ((1334*123.9)7(794+ 123.9)) = 183 ppm

8                Next, because RfCs are typically expressed in units of mg/m3, the HEC value in
9   ppm was converted using the conversion factor specific to methanol of 1 ppm =1.31 mg/m3:
                                             C-12
                        DRAFT - DO NOT CITE OR QUOTE

-------
                            HEC (mg/m3) =1.31 x 183 ppm = 240 mg/m3
 2          Finally, this HEC value was divided by a composite 100-fold UF (3 for uncertainty
 3    associated with animal to human differences, 10 for consideration of human variability, and 3 for
 4    database deficiencies) to obtain the chronic inhalation reference value:

 5                          RfC (mg/m3) = 240 mg/m3 - 100 = 2.4 mg/m3

      C.2. DECREASED BRAIN WEIGHT IN MALE RATS EXPOSED DURING
      GESTATION ONLY (GD7-GD17) (NEDO, 1987)
 6          Cmax, as calculated by the EPAs PBPK model, was selected as the dose metric for this
 7    exposure scenario, in concordance with the choice of this dose metric for the increased incidence
 8    of cervical rib in mice in the Rogers et al.  study (1993). Exposures occurred only during the
 9    major period of organogenesis in both studies. As there is evidence that  Cmaxis a better predictor
10    of response than  AUC for incidence of cervical rib (see Appendix D), it was assumed appropriate
11    to consider Cmax the better predictor for decreased brain weight as well.
12          The first  step in the current analysis is to convert the inhalation doses, given as ppm
13    values from the studies, to an internal dose surrogate or dose metric using the EPA PBPK model
14    (see Section 3.4). Predicted Cmax values for methanol in the blood of rats are  summarized in
15    Table C-4.

          Table C-4.  EPA's PBPK model estimates of methanol blood levels (Cmax) in rats
          following inhalation exposures
Exposure level (ppm)
200
1,000
5000
Methanol in lood Cmax (mg /L)a in
rats
1.2
10.6
630.5
     aCmax values were obtained by simulating 22 hr/day exposures

16          The current BMD technical guidance (U.S. EPA, 2000b) suggests that in the absence of
17   knowledge as to what level of response to consider adverse, a change in the mean equal to
18   1 control S.D. from the control mean can be used as a BMR for continuous endpoints. However,
19   it has been suggested that other BMRs, such as 5% change relative to estimated control mean,
20   are also appropriate when performing BMD analyses on fetal weight change as a developmental
21   endpoint (Kavlock et al., 1995).  Therefore, in this assessment, both a 1 control mean S.D.
22   change and a 5% change relative to estimated control mean were considered. All models were fit
23   using restrictions and option settings suggested in the EPAs BMD technical guidance document
24   (U.S. EPA, 2000b).

                                              C-13           DRAFT - DO NOT CITE OR QUOTE

-------
     C.2.1. BMD Approach with a BMR of 1 Control Mean S.D. (GD7-GD17) (NEDO, 1987)
 1          A summary of the results most relevant to the development of a POD using the BMD
 2   approach (BMD, BMDL, and model fit statistics) for decreased brain weight at 8 weeks in male
 3   rats exposed to methanol during gestation from days 7-17, with a BMR of 1 control mean S.D, is
 4   provided in Table C-5.  Male brain weight responses were chosen because they resulted in lower
 5   BMD and BMDL estimates than female responses (data not shown). Model fit was determined
 6   by  statistics (AIC and j^ residuals of individual dose groups) and visual inspection, as
 7   recommended by EPA (2000b).  The polynomial and power models reduced to linear form and
 8   returned identical modeling results. In contrast, the more complex Hill and Exponential
 9   models, which estimate a response "plateau" or asymptote, returned similar, markedly nonlinear
10   results. This is because these models approximated the response "plateau" to be near the
11   maximum drop in brain weight observed in the study (approximately 10% at the high dose),
12   resulting in a distinctly "L" shaped dose-response curve.89 In this case, the only PBPK model
13   estimated Cmax dose that is associated with a significant response over controls, the high-dose, is
14   60-fold higher than the mid-dose Cmax estimate. Thus, there are many plausible curve shapes
15   and,consequently, a wide range of BMDL estimates. Per EPA (2000b) guidance and to err on the
16   side of public  health protection, the lowest BMDLiso of 10.26 mg methanol/L in blood estimated
17   from adequate and plausible models was chosen for use in the RfC derivation. However, it
18   should be noted that there is a great deal of uncertainty and model dependence associated with
19   these dose-response data.

          Table C-5.  Comparison of BMDisD results for decreased brain weight in male rats at 8
          weeks of age using modeled Cmax of methanol as a dose metric
Model
Linear
2nd degree polynomial
3rd degree polynomial
Power
BMD1SD (Cmax,
mg/L)a
207.18
207.18
207.18
207.18
BMDL1SD (Cmax,
mg/L)a
135.22
135.22
135.22
135.22
/7-value
0.7881
0.7881
0.7881
0.7881
AIC
-173.12
-173.12
-173.12
-173.12
Scaled residuald
-0.43
-0.43
-0.43
-0.43
Hillb 43.08 10.26 0.9602 -171.59 -0.10
Exponential! 199.98 127.55 0.9494 -173.13 -0.42
Exponentials 199.98 127.55 0.9494 -173.13 -0.42
Exponential 4b 39.53 10.26 Not reported -171.59 0.10
     89 The extent of the "L" shape is dependent on the asymptote term, or "plateau" level, estimated for the data. If the
     asymptote term (v) in the Hill model is set to -.4 (representing a 20% drop from the control brain weight of 2
     grams), the model result is more linear and the BMD and BMDL estimates are approximately 4-fold higher.
                                               C-14            DRAFT - DO NOT CITE OR QUOTE

-------
      aThe BMDL is the 95% lower confidence limit on the Cmax estimated to decrease brain weight by 1 control mean
      S.D. using BMDS model options and restrictions suggested by EPA BMD technical guidance (U.S. EPA, 2000b).
      bPer EPA (2000b) guidance and to err on the side of public health protection, the lowest BMDL1SD of 10.26 mg
      methanol/L in blood estimated from adequate and plausible models was chosen for use in the RfC derivation
      °AIC = Akaike Information Criterion = -2L + 2P, where L is the log-likelihood at the maximum likelihood
      estimates for the parameters, and P is the number of modeled degrees of freedom (usually the number of parameters
      estimated).
      d% d residual (measure of how model-predicted responses deviate from the actual data) for the dose group  closest
      to the BMD scaled by an estimate of its S.D. Provides a comparative measure of model fit  near the BMD.
      Residuals that exceed 2.0 in absolute value should cause one to question model fit in this region.

      Source: NEDO (1987)

 1    ====================================================================
 2             Hill Model.  (Version:  2.14;  Date:  06/26/2008)
 3             Input Data  File: C:\Usepa\BMDS21\Data\Methanol\NEDO\Gest-only\hilm-8wk-
 4    brwHil-Restrict.(d)
 5             Gnuplot Plotting File:  C:\Usepa\BMDS21\Data\Methanol\NEDO\Gest-only\hilm-
 6    8wk-brwHil-Restrict.plt
 7                                                     Tue Aug  25  12:40:30 2009
 8     ====================================================================
 9
10     BMDS Model  Run
j j
12
13       The form  of the response function is:
14
15       Y[dose] = intercept + v*dose/xn/(k^n + dose^n)
16
17
18       Dependent variable = Mean
19       Independent variable = Dose
20       Power parameter restricted  to be  greater than 1
21       The variance is to be modeled as  Var(i)  = exp(lalpha  + rho * ln(mean(i)))
22
23       Total number of dose groups  = 4
24       Total number of records with  missing values  = 0
25       Maximum number  of iterations  = 250
26       Relative  Function Convergence has been set  to:  le-008
27       Parameter Convergence has been set to:  le-008
28
29
30
31                       Default Initial  Parameter Values
32                              lalpha  =     -4.68678
33                                 rho  =            0
34                           intercept  =            2
35                                   v  =        -0.19
36                                   n  =     0.861776
37                                    k  =      303.331
38
39
40             Asymptotic Correlation  Matrix of Parameter Estimates
41
42         ( *** The model  parameter(s)   -n
43              have been estimated  at a boundary point,  or have been specified by the user,
44              and do not appear in  the correlation  matrix )
45
46                     lalpha           rho    intercept             v             k
47
48        lalpha             1            -1       -0.083           0.6        -0.18
49
50           rho            -1             1        0.096          -0.6          0.18
51


                                                 C-15            DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
 intercept


         v


         k
    -0.083


       0.6


     -0.18
 0.096


  -0. 6


  0.18
    1


 0.19


-0.55
 0.19


    1


-0.73
-0.55


-0.73


    1
                            Parameter  Estimates
Variable
lalpha
rho
intercept
        v
        n
        k
    Estimate
  7.03732
 -18.1432
   2.0068
-0.232906
        1
  121.949
   Std.  Err.
  4.98399
  7.32604
0.0134454
0.0881494
       NA
  194.687
         95.0% Wald Confidence  Interval
   Lower Conf. Limit   Upper  Conf.  Limit
       -2.73112            16.8058
        -32.502            -3.78448
        1.98045            2.03316
      -0.405676          -0.0601362
                                                   -259.631
                                                                       503.529
NA - Indicates that this parameter has  hit  a  bound
     implied by some inequality constraint  and  thus
     has no standard error.
     Table of Data and Estimated Values  of  Interest


 Dose       N    Obs Mean     Est Mean   Obs  Std Dev  Est Std Dev   Scaled Res.
0
1.2
10.6
630.5
11
11
12
10
2
2.01
1.99
1.81
2.01
2
1.99
1.81
0.047
0.075
0.072
0.161
0.0608
0.0614
0.0662
0.154
-0.371
0.295
0.0954
-0.0338
 Model Descriptions for likelihoods  calculated
 Model Al:        Yij  = Mu(i)  + e(ij)
           Var{e(ij)}  = Sigma/x2


 Model A2:        Yij  = Mu(i)  + e(ij)
           Var{e(ij)}  = Sigma(i)^2


 Model A3:        Yij  = Mu(i)  + e(ij)
           Var{e(ij)}  = exp(lalpha  +  rho*ln(Mu(i)))
     Model A3 uses any fixed variance parameters  that
     were specified by the user


 Model  R:         Yi  = Mu + e(i)
            Var{e (i) }  = Sigma/x2
                       Likelihoods  of  Interest
            Model
             Al
             A2
             A3
         fitted
              R
Log (likelihood)
83.205960
92.060485
90.797178
90.795933
70.761857
# Param' s
5
8
6
5
2
AIC
-156.411920
-168.120970
-169.594356
-171.591867
-137.523714
                   Explanation of Tests
                                              C-16
                                                        DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
 Test 1:  Do responses and/or variances  differ  among Dose levels?
          (A2 vs.  R)
 Test 2:  Are Variances Homogeneous?  (Al vs A2)
 Test 3:  Are variances adeguately modeled?  (A2 vs. A3)
 Test 4:  Does the Model for the Mean Fit?  (A3  vs.  fitted)
 (Note:   When rho=0 the results  of Test  3 and Test  2 will be the same.)
   Test


   Test 1
   Test 2
   Test 3
   Test 4
                     Tests of Interest
-2*log(Likelihood Ratio)   Test  df
            42.5973
            17.7091
            2.52661
         0.00248896
   p-value


  <.0001
0.000505
  0.2827
  0.9602
The p-value for Test 1 is less  than .05.   There  appears to be a
difference between response and/or variances  among  the dose levels
It seems appropriate to model  the  data
The p-value for Test 2 is less  than .1.
model appears to be appropriate
                             A non-homogeneous variance
The p-value for Test 3 is greater than .1.
 to be appropriate here


The p-value for Test 4 is greater than .1.
to adeguately describe the data
                                 The modeled variance appears
                                 The model chosen seems
        Benchmark Dose Computation


Specified effect =             1


Risk Type        =     Estimated  standard  deviations from the control mean


Confidence level =           0.95


             BMD =        43.0842


            BMDL =       10.2551
                                              C-17
                                                        DRAFT - DO NOT CITE OR QUOTE

-------
                                      Hill Model with 0.95 Confidence Level
               2.05
      o
      Q.
      (/)
      0)
      o:
      c
      (0
      0)
                1.7
                                                                             600
        12:4008/252009
         Figure C-3. Hill model, BMR of 1 Control Mean S.D. - Decreased Brain weight in
         male rats at 8 weeks age versus Cmax, Gestation only inhalational study.
           Source: NEDO(1987).
1          Once the BMDLiso was obtained in units of mg/L, it was used to derive a chronic
2   inhalation reference value. The first step is to calculate the HEC using the PBPK model
3   described in Appendix B. An algebraic equation is provided (Equation 1 of Appendix B) that
4   describes the relationship between predicted methanol AUC and the human equivalent inhalation
5   exposure concentration (HEC) in ppm. This equation can also be used to estimate model
6   predictions for HECs from Cmax values because Cmax values and AUC values, were estimated at
7   steady-state for constant 24-hour exposures (i.e., AUC = 24 x Cmax).

8        BMDLHEC (ppm) = 0.0224*BMDLisD*24+(1334*BMDLisD*24)/(794+ BMDLiSD*24)
9         BMDLHEC (ppm) = 0.0224*10.3*24 + ((1334*10.3*24)7(794+ 10.3*24)) = 322 ppm
                                            C-18
DRAFT - DO NOT CITE OR QUOTE

-------
 1          Next, because RfCs are typically expressed in units of mg/m3, the HEC value in ppm was
 2   converted using the conversion factor specific to methanol of 1 ppm =1.31 mg/m3:

 3                          HEC (mg/m3) = 1.31 x 322 ppm = 422 mg/m3
3
 4          Finally, this HEC value was divided by a composite 100-fold UF (3 for uncertainty
 5   associated with animal to human differences, 10 for consideration of human variability, and 3 for
 6   database deficiencies) to obtain the chronic inhalation reference value:

 7                          RfC (mg/m3) = 422 mg/m3 - 100 = 4.2 mg/m3

     C.2.2. BMD Approach with a BMR of 0.05 Change Relative to Control Mean (GD7-GD17)
     (NEDO, 1987)
 8          A summary of the results most relevant to the development of a POD using the BMD
 9   approach (BMD, BMDL, and model fit statistics) for decreased brain weight at 8 weeks in male
10   rats exposed to methanol during gestation from days 7 to 17, with a BMR of 0.05 change relative
11   to estimated control mean, is provided in Table C-6. Model fit was determined by statistics (AIC
12   and ^ residuals of individual dose groups) and visual inspection, as recommended by EPA
13   (2000b). Modeling considerations and uncertainties for this dataset were discussed in C.2.1 and,
14   as was done for the BMR of 1 S.D., the lowest BMDLos of 21.07 mg methanol/L in blood
15   estimated from adequate and plausible models was chosen for use in the RfC derivation.
                                              C-19           DRAFT - DO NOT CITE OR QUOTE

-------
     Table C-6. Comparison of BMD0s modeling results for decreased brain weight in male
     rats at 8 weeks of age using modeled Cmax of methanol as a common dose metric
Model
Linearb
2nd degree polynomial
3rd degree polynomial
Power
BMD05 (Cmax,
mg/L)a
328.84
328.84
328.84
328.84
BMDL05 (Cmax,
mg/L)a
226.08
226.08
226.08
226.08
^-value
0.7881
0.7881
0.7881
0.9446
AICC
-173.12
-173.12
-173.12
-173.12
Scaled residual11
0.02
0.02
0.02
0.02
Hillb 92.30 Not reported 0.9602 -171.59 0.10
Exponential! 320.62 215.13 0.9494 -173.13 0.02
Exponentials 320.62 215.13 0.9494 -173.13 0.02
Exponential 4b 76.36 21.07 Not reported -171.59 0.10
aThe BMDL is the 95% lower confidence limit on the Cmax estimated to decrease brain weight by 5% using BMDS
model options and restrictions suggested by EPA BMD technical guidance (U.S. EPA, 2000b).
bPer EPA (2000b) guidance and to err on the side of public health protection, the lowest BMDL05 of 21.07 mg
methanol/L in blood estimated from adequate and plausible models was chosen for use in the RfC derivation.
°AIC = Akaike Information Criterion = -2L + 2P, where L is the log-likelihood at the maximum likelihood
estimates for the parameters, and P is the number of modeled degrees of freedom (usually the number of parameters
estimated).
dj^d residual (measure of how model-predicted responses deviate from the actual data) for the dose group closest to
the BMD scaled by an estimate of its S.D. Provides a comparative measure of model fit near the BMD. Residuals
that exceed 2.0 in absolute value should cause one to question model fit in this region.

Source:  NEDO (1987).
1
2
O
4
5
6
7
/
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
Exponential Model. (Version: 1.61; Date: 7/24/2009)
Input Data File: C:\Usepa\BMDS21\Data\Methanol\NEDO\Gest-only\expm-8wk
brwSetting. (d)
Gnuplot Plotting File:
Tue Aug 25 14:15:15 2009

BMDS Model
	

The form
Model
Model
Model
Model


Run
	



	 	



„ 	

of the response function by Model:
2:
3:
4:
5:

Note: Y[dose]
sign =
sign =

Model
Model
Model


Dependent

2 is
3 is
4 is


Y[dose] = a
Y[dose] = a
Yfdose] = a
Yfdose] = a

* expfsign * b * dose}
* expfsign * (b * dosej^d}
* [c-(c-l) * exp{-b * dose}]
* [c-(c-l) * exp{-(b * dose)Ad}]

is the median response for exposure = dose;
+1 for increasing trend in data;
-1 for decreasing trend.

nested within
nested within
nested within



Models 3 and 4.
Model 5 .
Model 5.


variable = Mean
Independent variable = Dose
Data are
Variance
assumed to be distributed: normally
Model
: exp(lnalpha
+rho *ln(Y[dose] ) )
The variance is to be modeled as Var(i) = exp(lalpha + log(mean(i)) * rho)
                                              C-20
DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
Total number of dose groups  =  4
Total number of records  with missing values = 0
Maximum number of iterations = 250
Relative Function Convergence  has been set to: le-008
Parameter Convergence has  been set to: le-008


MLE solution provided: Exact
  Variable


   Inalpha
       rho
         a
         b
         c
         d
          Initial Parameter Values


Model 2           Model  3           Model 4
      NC
      NC
      NC
      NC
      NC
      NC
NC

NC

NC

NC

NC

NC
   7.32457
  -18.5236
    2.1105
0.00239093
  0.816778
  Variable


   Inalpha
       rho
         a
         b
         c
         d
        Parameter Estimates  by Model


Model 2           Model  3           Model 4
   NC
   NC
   NC
   NC
   NC
   NC
NC

NC

NC

NC

NC

NC
   7.03418
  -18.1386
   2.00677
0.00941775
  0.902498
  NC = No Convergence



         Table of Stats  From Input Data


  Dose      N         Obs  Mean     Obs Std Dev
0
1.2
10. 6
630.5

Model
4



11
11
12
10

Do:

1,
10,
630,





36
0
.2
.6
.5
2
2.01
1.99
1.81
Estimated
Est
2
2
1
1




Values
Mean
.007
.005
.988
.812
0
0
0
0






.047
.075
.072
.161
of Interest
Est Std
0.06082
0.06142
0.06617
0.1538





Scaled Residual
-0.3692
0.2932
0.09527
-0.03335
Model 5


  NC
  NC
  NC
  NC
  NC
  NC
Model 5


  NC

  NC

  NC

  NC

  NC

  NC
Other models for which likelihoods are calculated:


  Model Al:         Yij  =  Mu(i)  +  e(ij)
            Var{e(ij)}  =  SigmaA2


  Model A2:         Yij  =  Mu(i)  +  e(ij)
            Var{e(ij)}  =  Sigma(i)^2
          Model A3:
                   Yij  =  Mu(i)  +  e(ij)
                                              C-21
                                                     DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
            Var{e(ij)}  = exp(lalpha + log(mean(i)


  Model  R:        Yij  = Mu + e(i)
            Var{e(ij)}  = Signal
                                    rho)
                  Model
             Likelihoods of  Interest


             Log (likelihood)      DF
                                                            AIC
Al
A2
A3
R
4
83.20596
92.06049
90.61606
70.76186
90.79579
5
8
6
2
5
-156.4119
-168.121
-169.2321
-137.5237
-171.5916
Additive constant for all log-likelihoods  =      -40.43.  This constant added to the
above values gives the log-likelihood including  the  term that does not
depend on the model parameters.



                              Explanation  of  Tests


Test 1:  Does response and/or variances  differ among Dose  levels?  (A2 vs. R)
Test 2:  Are Variances Homogeneous?  (A2  vs. Al)
Test 3:  Are variances adeguately modeled? (A2 vs. A3)


Test 6a: Does Model 4 fit the data?  (A3  vs 4)
  Test


  Test 1
  Test 2
  Test 3
 Test 6a
         Tests  of  Interest


-2*log(Likelihood  Ratio)


                  42.6
                17.71
                2.889
               -0.3595
D. F.


  6
  3
  2
  1
p-value
 < 0.0001
 0.000505
   0.2359
      N/A
  The p-value for Test 1 is less  than .05.   There  appears  to be a
  difference between response and/or variances  among  the dose
  levels, it seems appropriate to model  the  data.


  The p-value for Test 2 is less  than .1.  A non-homogeneous
  variance model appears to be appropriate.


  The p-value for Test 3 is greater than .1.  The  modeled
  variance appears to be appropriate here.


  The p-value for Test 6a is less than .1.   Model  4 may not adeguately
  describe the data; you may want to consider another model.
Benchmark Dose Computations:


  Specified Effect = 0.050000


         Risk Type = Relative deviation


  Confidence Level = 0.950000
           Model
             BMD and BMDL by Model


                     BMD
                                                BMDL


                                              C-22
                                                     DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
               o
               o
         76.3561
               0
0
0
21.0664
0
Not computed
Not computed

Not computed
         o
         Q.
                 2.05
                 1.95
1.9
                 1.85
                  1.8
                 1.75
                  1.7
                                   Exponential Model 4 with 0.95 Confidence Level
                                 Exponential
                      BMDL
              BMD
                        0
               100
200
300
 dose
400
500
600
          14:1508/252009
          Figure C-4. Exponential model, BMR of 0.05 relative risk - Decreased Brain weight
          in male rats at 8 weeks age versus Cmax, Gestation only inhalational study.
            Source: NEDO(1987).

 6          Once the BMDLos was obtained in units of mg/L, it was used to derive a chronic
 7   inhalation reference value.  The first step is to calculate the HEC using the PBPK model
 8   described in Appendix B. An algebraic equation is provided (Equation 1 of Appendix B) that
 9   describes the relationship between predicted methanol AUC and the human equivalent inhalation
10   exposure concentration (HEC) in ppm. This equation can also be used to estimate model
11   predictions for HECs from  Cmax values because Cmax values, and AUC values were estimated at
12   steady-state for constant 24-hour exposures (i.e., AUC = 24 x Cmax).

13         BMDLHEC (ppm) = 0.0224*BMDL05*24+(1334*BMDLo5*24)/(794+ BMDL05*24)
14        BMDLHEC (ppm) = 0.0224*21.07*24 + ((1334*21.07*24)7(794+ 21.07*24)) = 530 ppm
                                              C-23
                                           DRAFT - DO NOT CITE OR QUOTE

-------
 1          Next, because RfCs are typically expressed in units of mg/m3, the HEC value in ppm was
 2    converted using the conversion factor specific to methanol of 1 ppm =1.31 mg/m3:
                            HEC (mg/m3) =1.31 x 530 ppm = 695 mg/m3
 4          Finally, this HEC value was divided by a composite 100-fold UF (3 for uncertainty
 5    associated with animal to human differences, 10 for consideration of human variability, and 3 for
 6    database deficiencies) to obtain the chronic inhalation reference value:
                                      3
                                                   3
 7                          RfC (mg/m") = 695 mg/m" - 100 = 7.0 mg/m"

      C.3. RfC DERIVATIONS USING ROGERS ET AL. (1993)
 8          For the purposes of deriving an RfC for methanol from developmental endpoints using
 9    the BMD method and mouse data, cervical rib incidence data were evaluated from Rogers et al.
10    (1993). In this paper, Rogers et al. (1993) also utilized a BMD methodology, examining the
11    dosimetric threshold for cervical ribs and other developmental impacts by applying a log-logistic
12    maximum likelihood model to the dose-response data. Using air exposure concentrations (ppm)
13    as their dose metric, a value for the lower 95% confidence limit on the benchmark dose for 5%
14    additional risk in mice was 305 ppm (400 mg/m3), using the log-logistic model.  Although the
15    teratology portion of the NEDO study (1987) also reported increases in cervical rib incidence in
16    Sprague-Dawley rats, the Rogers et al. (1993) study was chosen for dose-response modeling
17    because effects were seen at lower doses, it was peer-reviewed and published in the open
18    literature, and data on individual animals were available for a more statistically robust analysis
19    utilizing nested models available in BMDS.
20          The first step in the current BMD analysis is to convert the inhalation doses, given as
21    ppm values from the studies,  to an internal dose surrogate or dose metric using the EPA's PBPK
22    model (see Section 3.4). For cervical rib malformations, Cmax of methanol in blood (mg/L) is
23    chosen as the appropriate internal dose metric metric (see Appendix D for further explanation).
24    Predicted Cmax values for methanol in the blood of mice are summarized in Table C-7.

          Table C-7. EPA's PBPK model estimates of methanol blood levels (Cmax) in mice
          following inhalation exposures
Exposure concentration (ppm)
1
10
50
100
250
Methanol in blood Cmax (mg/L)a
in mice
0.0216
0.218
1.14
2.46
7.83
                                              C-24
DRAFT - DO NOT CITE OR QUOTE

-------
Exposure concentration (ppm)
500
1,000
Methanol in blood Cmax (mg/L)a in mice
26.4
134
      aRounded to three significant figures.
 1           These Cmax values are then used as the dose metric for the BMD analysis of cervical rib
 2    incidence.
 3           A 10% BMR level is the value typically calculated for comparisons across chemicals and
 4    endpoints for dichotomous responses because this level is near the low end of the observable
 5    range for many types of toxicity studies.  However, reproductive and developmental studies
 6    having a nested design often have a greater sensitivity, and a 5% BMR is typically appropriate
 7    for determination of a POD (U.S. EPA, 2000b; Allen et al., 1994). Rogers et al. (1993) utilized a
 8    5% added risk for the BMR in the original study. This assessment utilizes both a 10% and 5%
 9    extra risk level as a BMR for the determination of a POD.90 The nested suite of models available
10    in BMDS was used to model the cervical rib data. In general, data from developmental toxicity
11    studies are best modeled using nested models, as these models account for any intralitter
12    correlation (i.e., the tendency of littermates to respond similarly to one another relative to other
13    litters in a dose group).  All models were fit using restrictions and option settings suggested in
14    the EPAs BMD technical guidance document (U.S. EPA, 2000b).

      C.3.1. BMD Approach with a BMR of 0.10 Extra Risk (Rogers et al., 1993)
15           A summary of the results most relevant to the development of a POD using the BMD
16    approach (BMD, BMDL, and model fit statistics) for increased incidence of cervical rib in mice
17    exposed to methanol during gestation from days 6 to 15, with a BMR of 0.10 extra risk, is
18    provided in Table C-8.  Model fit was determined by statistics (AIC and $ residuals of individual
19    dose groups) and visual inspection, as recommended by U.S. EPA (2000b). The best model fit to
20    these data (from visual inspection and comparison of AIC values) was obtained using the Nested
21    Logistic (NLogistic) model. The BMDLio was determined to be 94.3 mg/L using the 95% lower
22    confidence limit of the dose-response curve expressed in terms of the Cmax for methanol in blood.

          Table C-8. Comparison of BMD modeling results for cervical rib incidence in mice
          using modeled Cmax of methanol as a common dose metric
Model
NLogisticb
NCTR
BMD10
(Cmax, mg/L)a
141.492
207.945
BMDL10 (Cmax,
mg/L)a
94.264
103.972
/7-value
0.293
0.241
AIC
1046.84
1048.92
Scaled residual"1
0.649
0.662
      90 Starr and Festa (2003) have argued that the Rogers et al. (1993) study's experimental design lacked the statistical
      power to detect a 5% risk and that a 5% level lay below the observable response data. However, EPA's BMD
      guidance (U.S. EPA, 2000b) does not preclude the use of a BMR that is below observable response data and EPA
      has deemed that the Rogers et al. (1993) is adequate for the consideration of a 5% BMR.
                                                C-25
DRAFT - DO NOT CITE OR QUOTE

-------
Model
Rai and Van Ryzin
BMD10
(Cmax, mg/L)a
221.509
BMDL10 (Cmax,
mg/L)a
110.754
/7-value
0.163
AICC
1051.65
Scaled residuald
0.661
      aDaily Cmax was estimated using a mouse PBPK model as described in section 3.4 of the methanol lexicological
      review; the BMDL is the 95% lower confidence limit on the Cmax for a 10% extra risk (dichotomous endpoints)
      estimated by the model using the likelihood profile method (U.S. EPA, 2000b).
      bModel choice based on adequate p value (> 0.1), visual inspection, low AIC, and low (absolute) scaled residual.
      °AIC = Akaike Information Criterion = -2L + 2P, where L is the log-likelihood at the maximum likelihood
      estimates for the parameters, and P is the number of modeled degrees of freedom (usually the number of parameters
      estimated).
      dj^d residual (measure of how model-predicted responses deviate from the actual data) for the dose group closest to
      the BMD scaled by an estimate of its S.D. Provides a comparative measure of model fit near the BMD. Residuals
      exceeding 2.0 in absolute value should cause one to question model fit in this region.

      Source:  Rogers etal. (1993).

 1    ====================================================================
 2    NLogistic Model.
 3    (Version: 2.13;  Date:   02/20/2007)
 4    Input Data  File:  U:\Methanol\BMDS\CervicalRib\Cmax\NLog_Cmax_10_default.(d)
 5    Wed Nov  07  15:45:40  2007
 6    ====================================================================
 7
 8    BMD Method  for RfC:  Incidence  of Cervical Rib in Mice versus Cmax Methanol, GD 6-15
 9    inhalational study (Rogers,et  al.,  1993)
JQ
11      The probability function  is:
12
13     Prob. = alpha +  thetal*Rij +  [1 -  alpha - thetal*Rij]/
14
15                           [l+exp(-beta-theta2*Rij-rho*log(Dose) ) ] ,
16
17              where  Rij  is the  litter specific covariate.
18
19     Restrict Power  rho  >= 1.
20
21     Total number  of  observations  =  166
22     Total number  of  records with  missing values  =  0
23     Total number  of  parameters in model = 9
24     Total number  of  specified  parameters = 0
25
26     Maximum number  of iterations  =  250
27     Relative Function Convergence has  been set  to:  le-008
28     Parameter  Convergence has  been  set to:  le-008
29
30                       Default Initial Parameter Values
31                               alpha  =     0.297863
32                                 beta  =     -7.94313
33                              thetal  =             0
34                              theta2  =             0
35                                 rho  =      1.09876
36                                 phil  =     0.213134
37                                 phi2  =     0.309556
38                                 phi3  =     0.220142
39                                 phi4  =     0.370587
40
41                                      Parameter Estimates
42
43           Variable            Estimate               Std.  Err.
44              alpha              0.102434              *
45                beta              -4.80338              *
46              thetal            0.0325457              *
47              theta2            -0.436115              *

                                                 C-26            DRAFT - DO NOT CITE OR QUOTE

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
rho 1 *
phil 0.200733 *
phi2 0.307656 *
phi3 0.212754 *
phi4 0.368426 *
* - Indicates that this value is not calculated.
Log-likelihood: -515.422 AIC: 1046.84
Litter Data
Lit. -Spec. Litter
Dose Cov. Est. Prob. Size Expected Observed

0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.

,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000

1.
1.
2.
2.
2.
2.
2.
3.
3.
3.
3.
4.
4.
4.
4.
5.
5.
5.
5.
5.
5.
5.
5.
5.
5.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
7.
7.
7.
7.
7.
7.
7.
7.
7.
7.
7.

0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000

0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

.135
.135
.168
.168
.168
.168
.168
.200
.200
.200
.200
.233
.233
.233
.233
.265
.265
.265
.265
.265
.265
.265
.265
.265
.265
.298
.298
.298
.298
.298
.298
.298
.298
.298
.298
.298
.298
.298
.298
.298
.298
.330
.330
.330
.330
.330
.330
.330
.330
.330
.330
.330

1
1
2
2
2
2
2
3
3
3
3
4
4
4
4
5
5
5
5
5
5
5
5
5
5
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
7
7
7
7
7
7
7
7
7
7
7

0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.

,135
,135
,335
,335
,335
,335
,335
,600
, 600
, 600
,600
,930
,930
,930
,930
,326
,326
,326
,326
,326
,326
,326
,326
,326
,326
,786
,786
,786
,786
,786
,786
,786
,786
,786
,786
,786
,786
,786
,786
,786
,786
,312
,312
,312
,312
,312
,312
,312
,312
,312
,312
,312

0
0
0
1
0
2
0
0
0
1
1
0
1
0
1
0
1
3
1
0
1
1
0
0
1
3
6
1
0
0
1
2
0
2
3
5
0
3
3
3
5
0
1
2
3
2
3
5
0
2
5
1
Scaled
Residual

-0.3950
-0.3950
-0.5790
1.1490
-0.5790
2.8770
-0.5790
-0.7317
-0.7317
0.4874
0.4874
-0.8699
0.0650
-0.8699
0.0650
-1.0004
-0.2458
1.2632
-0.2458
-1.0004
-0.2458
-0.2458
-1.0004
-1.0004
-0.2458
0.7656
2.6578
-0.4959
-1.1267
-1.1267
-0.4959
0.1348
-1.1267
0.1348
0.7656
2.0271
-1.1267
0.7656
0.7656
0.7656
2.0271
-1.2513
-0.7100
-0.1688
0.3725
-0.1688
0.3725
1.4551
-1.2513
-0.1688
1.4551
-0.7100
C-27
DRAFT - DO NOT CITE OR QUOTE

-------
1
2
3
4
5
6
7
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
0.
0.
0.
0.
0.
0.
0.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.

526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000

,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
7.
7.
8.
8.
8.
8.
8.
1.
1.
2.
2.
3.
3.
3.
3.
4.
4.
5.
5.
5.
5.
5.
6.
6.
7.
7.
7.
7.
7.
7.
8.
8.
8.

2.
3.
4.
4.
5.
5.
5.
5.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
7.
7.
7.
7.
7.
7.
7.
7.
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000

0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
.330
.330
.363
.363
.363
.363
.363
.494
.494
.430
.430
.383
.383
.383
.383
.356
.356
.346
.346
.346
.346
.346
.350
.350
.363
.363
.363
.363
.363
.363
.383
.383
.383

.703
. 631
.562
.562
.506
.506
.506
.506
.466
.466
.466
.466
.466
.466
.466
.466
.466
.466
.466
.466
.466
.466
.466
.466
.444
.444
.444
.444
.444
.444
.444
.444
7
7
8
8
8
8
8
1
1
2
2
3
3
3
3
4
4
5
5
5
5
5
6
6
7
7
7
7
7
7
8
8
8

2
3
4
4
5
5
5
5
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
7
7
7
7
7
7
7
7
2.
2.
2.
2.
2.
2.
2.
0.
0.
0.
0.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
2.
2.
2.
2.
2.
2.
2.
2.
3.
3.
3.

1.
1.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
3.
3.
3.
3.
3.
3.
3.
3.
,312
,312
,902
,902
,902
,902
,902
,494
,494
,859
,859
,150
,150
,150
,150
,425
,425
,732
,732
,732
,732
,732
,099
,099
,543
,543
,543
,543
,543
,543
,068
,068
,068

,406
,892
,250
,250
,530
,530
,530
,530
,796
,796
,796
,796
,796
,796
,796
,796
,796
,796
,796
,796
,796
,796
,796
,796
,105
,105
,105
,105
,105
,105
,105
,105
2
1
1
4
3
8
2
0
0
0
2
3
1
2
1
3
0
0
4
0
1
0
3
2
3
2
2
2
2
0
2
0
8

2
3
2
1
3
5
3
1
3
3
3
5
6
5
2
0
2
0
5
4
3
2
4
2
0
4
5
1
4
1
5
3
-0.1688
-0.7100
-0.9020
0.5204
0.0463
2.4170
-0.4279
-0.9887
-0.9887
-1.0732
1.4251
1.7287
-0.1400
0.7944
-0.1400
1.1858
-1.0729
-1.0898
1.4275
-1.0898
-0.4604
-1.0898
0.4839
-0.0534
0.2128
-0.2530
-0.2530
-0.2530
-0.2530
-1.1847
-0.4373
-1.2562
2.0195

0.8346
1.1101
-0.1967
-0.9842
0.3091
1.6241
0.3091
-1.0058
0.1162
0.1162
0.1162
1.2556
1.8253
1.2556
-0.4534
-1.5928
-0.4534
-1.5928
1.2556
0.6859
0.1162
-0.4534
0.6859
-0.4534
-1.5658
0.4511
0.9554
-1.0615
0.4511
-1.0615
0.9554
-0.0531
C-28
DRAFT - DO NOT CITE OR QUOTE

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
J ~s
60
61
62
63
64
65
66
67
526
526
526
526
526
526
526
526
526

2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005

.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000

.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000

7.
7.
7.
7.
8.
8.
8.
9.
9.

1.
1.
1.
2.
2.
3.
3.
3.
3.
4.
4.
4.
4.
4.
4.
4.
5.
5.
5.
5.
5.
5.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
7.
7.
7.
7.
7.
8.

Combine litters
within dose
the

fit of

0000
0000
0000
0000
0000
0000
0000
0000
0000

0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000

0.444
0.444
0.444
0.444
0.437
0.437
0.437
0.443
0.443

0.926
0.926
0.926
0.894
0.894
0.853
0.853
0.853
0.853
0.802
0.802
0.802
0.802
0.802
0.802
0.802
0.743
0.743
0.743
0.743
0.743
0.743
0. 681
0.681
0.681
0. 681
0. 681
0.681
0.681
0. 681
0. 681
0.681
0.681
0. 681
0. 623
0. 623
0.623
0.623
0. 623
0.576

with adjacent
groups until
the

7
7
7
7
8
8
8
9
9

1
1
1
2
2
3
3
3
3
4
4
4
4
4
4
4
5
5
5
5
5
5
6
6
6
6
6
6
6
6
6
6
6
6
7
7
7
7
7
8

levels of the
the expected count
3.
3.
3.
3.
3.
3.
3.
3.
3.

0.
0.
0.
1.
1.
2.
2.
2.
2.
3.
3.
3.
3.
3.
3.
3.
3.
3.
3.
3.
3.
3.
4.
4.
4.
4.
4.
4.
4.
4.
4.
4.
4.
4.
4.
4.
4.
4.
4.
4.

,105
,105
,105
,105
,496
,496
,496
,985
,985

,926
,926
,926
,789
,789
,559
,559
,559
,559
,208
,208
,208
,208
,208
,208
,208
,714
,714
,714
,714
,714
,714
,086
,086
,086
,086
,086
,086
,086
,086
,086
,086
,086
,086
,361
,361
,361
,361
,361
, 606

4
1
3
3
0
7
5
0
6

1
1
1
1
2
3
1
1
3
4
4
4
2
3
4
4
1
3
5
5
4
4
6
2
4
5
6
5
4
5
3
6
0
0
7
5
5
7
6
0

0
-1
-0
-0
-1
1
0
-1
0

0
0
0
-1
0
0
-1
-1
0
0
0
0
-1
-0
0
0
-1
-0
0
0
0
0
0
-1
-0
0
0
0
-0
0
-0
0
-2
-2
1
0
0
1
0
-1

.4511
.0615
.0531
.0531
.5793
.5832
.6796
.6270
.8225

.2834
.2834
.2834
.5502
.4157
.5454
.9294
.9294
.5454
.6851
.6851
.6851
.0440
.1795
.6851
.6851
.7660
.4648
.8364
.8364
.1858
.1858
.9945
.0836
.0445
.4750
.9945
.4750
.0445
.4750
.5641
.9945
.1227
.1227
.1486
.2781
.2781
.1486
.7133
.7419

litter-specific covariate
exceeds 3
.0, to help
improve
X^2 statistic to chi-square.








Grouped Data



0
0
0
0
0
0
0
0

Dose

.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000

Lit









Mean

.-Spec. Cov.

1.0000
2.0000
3.0000
4.0000
5.0000
5.0000
5.0000
5.0000












Expected Observed

0.270
1. 675
2.401
3.722
3.977
3.977
3.977
1.326










0
3
2
2
4
2
1
1
Scaled
Residual

-0.5586
1.0237
-0.2443
-0.8049
0.0098
-0.8614
-1.2970
-0.2458






















C-29
DRAFT - DO NOT CITE OR QUOTE

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.

134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.

526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.

2005.
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000

,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000

,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000

,0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
8.0000
8.0000
8.0000

1.0000
2.0000
3.0000
3.0000
4.0000
5.0000
5.0000
5.0000
6.0000
7.0000
7.0000
7.0000
8.0000
8.0000
8.0000

2.0000
3.0000
4.0000
5.0000
5.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
8.0000
8.0000
8.0000
9.0000
9.0000

1.0000
3.573
3.573
3.573
3.573
3.573
3.573
3.573
3.573
4.624
4.624
4. 624
4. 624
4.624
4.624
2.312
5.805
5.805
2.902
0.989
1.718
3.449
1.150
2.850
3.463
3.463
1.732
4.199
5.086
5.086
5.086
3.068
3.068
3.068
1.406
1.892
4.500
5.060
5.060
5.592
5.592
5.592
5.592
5.592
5.592
5.592
5.592
3.105
3.105
3.105
3.105
3.105
3.105
3.105
3.105
3.105
3.105
3.105
3.105
3.496
3.496
3.496
3.985
3.985
9
1
1
2
5
5
6
8
1
5
5
5
7
3
1
5
11
2
0
2
6
1
3
4
1
0
5
5
4
2
2
0
8
2
3
3
8
4
6
8
11
2
2
9
5
6
0
4
5
1
4
1
5
3
4
1
3
3
0
7
5
0
6
2
-1
-1
-0
0
0
1
1
-1
0
0
0
0
-0
-0
-0
1
-0
-1
0
1
-0
0
0
-1
-1
0
-0
-0
-1
-0
-1
2
0
1
-0
1
-0
0
0
2
-1
-1
1
-0
0
-1
0
0
-1
0
-1
0
-0
0
-1
-0
-0
-1
1
0
-1
0
.4207
.1474
.1474
.7013
.6367
.6367
.0827
.9747
.3869
.1441
.1441
.1441
.9096
.6214
.7100
.2698
.7418
.4279
.3982
.2488
.3759
.1400
.0799
.2388
.0962
.0898
.3044
.0284
.3578
.0166
.4373
.2562
.0195
.8346
.1101
.8351
.3670
.4926
.1644
.9700
.1785
.4469
.4469
.3729
.2384
.1644
.5658
.4511
.9554
.0615
.4511
.0615
.9554
.0531
.4511
.0615
.0531
.0531
.5793
.5832
.6796
.6270
.8225
2.777
     C-30
0.4909
DRAFT - DO NOT CITE OR QUOTE

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000
2005.0000

Chi-square =

To calculate
at the mean

2.0000
3.0000
3.0000
4.0000
4.0000
4.0000
4.0000
4.0000
4.0000
4.0000
5.0000
5.0000
5.0000
5.0000
5.0000
5.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
7.0000
7.0000
7.0000
7.0000
7.0000
8.0000

105.13 DF

the BMD and BMDL
litter specific

3.577
5.118
5.118
3.208
3.208
3.208
3.208
3.208
3.208
3.208
3.714
3.714
3.714
3.714
3.714
3.714
4.086
4.086
4.086
4.086
4.086
4.086
4.086
4.086
4.086
4.086
4.086
4.086
4.361
4.361
4.361
4.361
4.361
4.606

3
4
4
4
4
4
2
3
4
4
1
3
5
5
4
4
6
2
4
5
6
5
4
5
3
6
0
0
7
5
5
7
6
0

= 98 P-value = 0.

, the litter
covariate of


specific
all the

-0.8022
-0.9786
-0.9786
0.6851
0.6851
0.6851
-1.0440
-0.1795
0.6851
0.6851
-1.7660
-0.4648
0.8364
0.8364
0.1858
0.1858
0.9945
-1.0836
-0.0445
0.4750
0.9945
0.4750
-0.0445
0.4750
-0.5641
0.9945
-2.1227
-2.1227
1.1486
0.2781
0.2781
1.1486
0.7133
-1.7419

2930

covariate is fixed
data: 5.379518

C-31
DRAFT - DO NOT CITE OR QUOTE

-------
     Specified effect =
     Risk Type
     Confidence level =
                  BMD =
                 BMDL =
                                  0.1
                            Extra risk
                                 0.95
                              141.492
                              94.264
                                    Nested Logistic Model with 0.95 Confidence Level
                I
                      0.8
                      0.7
                      0.6
                      0.5
                      0.4
                      0.3
                      0.2
                 13:20 11/132007
                                                                          2000
          Figure C-5. Nested Logistic Model, 0.1 Extra Risk - Incidence of Cervical Rib in Mice
          versus Cmax Methanol, GD 6-15 inhalational study.
            Source:  Rogers et al. (1993).

 1          Once the BMDLio was obtained in units of mg/L, it was used to derive a chronic
 2   inhalation reference value.  The first step is to calculate the HEC using the PBPK model
 3   described in Appendix B. An algebraic  equation is provided (Equation 1 of Appendix B) that
 4   describes the relationship between predicted methanol AUC and the human equivalent inhalation
 5   exposure concentration (HEC) in ppm.  This equation can also be used to estimate model
 6   predictions for HECs from  Cmax values  because Cmax values and AUC values were estimated at
 7   steady-state for constant 24-hour exposures (i.e., AUC  = 24 x Cmax).
8
9
           BMDLHEC (ppm) = 0.0224*BMDLi0*24+(1334*BMDLi0*24)/(794+ BMDLi0*24)
          BMDLHEC (ppm) = 0.0224*94.3*24 + ((1334*94.3*24)7(794+ 94.3*24)) = 1038 ppm
10          Next, because RfCs are typically expressed in units of mg/m3, the HEC value in ppm was
1 1   converted using the conversion factor specific to methanol of 1 ppm =1.31 mg/m3:
12
                          HEC (mg/m3) = 1.31 x 1038 ppm = 1360 mg/m3
                                              C-32
                                                            DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
13
14
15
16
17
18
19
       Finally, this HEC value was divided by a composite 100-fold UF (3 for uncertainty
associated with animal to human differences, 10 for consideration of human variability, and 3 for
database deficiencies) to obtain the chronic inhalation reference value:

                       RfC (mg/m3) =1360 mg/m3 - 100 = 13.6 mg/m3

C.3.2. HMD Approach with a BMR of 0.05 Extra  Risk (Rogers et al., 1993)
       A summary of the results most relevant to the development of a POD using the BMD
approach (BMD, BMDL, and model fit statistics) for increased incidence of cervical rib in mice
exposed to methanol during gestation from days 6 to 15, with a BMR of 0.05 extra risk, is
provided in Table C-9. Model fit was determined by statistics (AIC and $ residuals of individual
dose groups) and visual inspection, as recommended by U.S. EPA (2000b). The best model fit to
these data (from visual inspection and comparison of AIC values) was obtained using the
NLogistic model.  The BMDL05 was determined to be 44.7 mg/L  using the 95% lower
confidence limit of the dose-response curve expressed in terms of the Cmaxfor methanol in blood.

     Table C-9. Comparison of BMD modeling results for cervical rib incidence in mice
     using modeled Cmax of methanol as a common dose metric
Model
NLogisticb
NCTR
Rai and Van
Ryzin
BMDOS
(Cmax,mg/L)a
67.022
101.235
107.838
BMDLos
(Cmax,mg/L)a
44.651
50.618
53.919
/7-value
0.293
0.241
0.163
AICC
1046.84
1048.92
1051.65
Scaled residual"1
0.649
0.662
0.661
           aDaily Cma^ was estimated using a mouse PBPK model as described in section 3.4 of the methanol
           toxicological review; the BMDL is the 95% lower confidence limit on the Cmax for a 5% extra risk
           (dichotomous endpoints) estimated by the model using the likelihood profile method (U.S. EPA,
           2000b).
           bModel choice based on adequate p value (> 0.1), visual inspection, low AIC, and low (absolute)
           scaled residual.
           °AIC = Akaike Information Criterion = -2L + 2P, where L is the log-likelihood at the maximum
           likelihood estimates for the parameters, and P is the number of modeled degrees of freedom (usually
           the number of parameters estimated).
           Yd residual (measure of how model-predicted responses deviate from the actual data) for the dose
           group closest to the BMD scaled by an estimate of its S.D. Provides a comparative measure of model
           fit near the BMD. Residuals exceeding 2.0 in absolute value should cause one to question model fit
           in this region.
           Source: Rogers et al. (1993).
NLogistic  Model.
 (Version:  2.13;  Date: 02/20/2007)
Input Data File: U:\Methanol\BMDS\CervicalRib\Cmax\NLog_Cmax_10_default.(d)
Wed Nov  07 15:45:40 2007
                                                 C-33
                                                            DRAFT - DO NOT CITE OR QUOTE

-------
BMD Method for RfC: Incidence of Cervical Rib in Mice versus Cmax Methanol, GD 6-15
inhalational  study  (Rogers, et al.,  1993)
J
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
55
56
57
58
59
60
61
62
63
64
65
66
The probability function is:

Prob. = alpha + thetal*Rij + [1 - alpha - thetal*Rij]/

[1+exp (-beta-theta2*Rij-rho*log(Dose) ) ] ,

where Rij is the litter specific covariate.

Restrict Power rho >= 1.

Total number of observations =166
Total number of records with missing values = 0
Total number of parameters in model = 9
Total number of specified parameters = 0

Maximum number of iterations = 250
Relative Function Convergence has been set to: le-008
Parameter Convergence has been set to: le-008

Default Initial Parameter Values
alpha = 0.297863
beta = -7.94313
thetal = 0
theta2 = 0
rho = 1.09876
phil = 0.213134
phi2 = 0.309556
phi3 = 0.220142
phi4 = 0.370587

Parameter Estimates

Variable Estimate Std. Err.
alpha 0.102434 *
beta -4.80338 *
thetal 0.0325457 *
theta2 -0.436115 *
rho 1 *
phil 0.200733 *
phi2 0.307656 *
phi3 0.212754 *
phi4 0.368426 *

* - Indicates that this value is not calculated.

Log-likelihood: -515.422 AIC: 1046.84
Litter Data

Lit. -Spec. Litter
Dose Cov. Est. Prob. Size Expected Observed
0.0000 1.0000 0.135 1 0.135 0
0.0000 1.0000 0.135 1 0.135 0
0.0000 2.0000 0.168 2 0.335 0
0.0000 2.0000 0.168 2 0.335 1
0.0000 2.0000 0.168 2 0.335 0
0.0000 2.0000 0.168 2 0.335 2
0.0000 2.0000 0.168 2 0.335 0
0.0000 3.0000 0.200 3 0.600 0
0.0000 3.0000 0.200 3 0.600 0
0.0000 3.0000 0.200 3 0.600 1
0.0000 3.0000 0.200 3 0.600 1
0.0000 4.0000 0.233 4 0.930 0
















































Scaled
Residual
-0.3950
-0.3950
-0.5790
1.1490
-0.5790
2.8770
-0.5790
-0.7317
-0.7317
0.4874
0.4874
-0.8699
                                        C-34
DRAFT - DO NOT CITE OR QUOTE

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.

134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
134.
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000

,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
4.
4.
4.
5.
5.
5.
5.
5.
5.
5.
5.
5.
5.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
7.
7.
7.
7.
7.
7.
7.
7.
7.
7.
7.
7.
7.
8.
8.
8.
8.
8.

1.
1.
2.
2.
3.
3.
3.
3.
4.
4.
5.
5.
5.
5.
5.
6.
6.
7.
7.
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000

0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
.233
.233
.233
.265
.265
.265
.265
.265
.265
.265
.265
.265
.265
.298
.298
.298
.298
.298
.298
.298
.298
.298
.298
.298
.298
.298
.298
.298
.298
.330
.330
.330
.330
.330
.330
.330
.330
.330
.330
.330
.330
.330
.363
.363
.363
.363
.363

.494
.494
.430
.430
.383
.383
.383
.383
.356
.356
.346
.346
.346
.346
.346
.350
.350
.363
.363
4
4
4
5
5
5
5
5
5
5
5
5
5
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
7
7
7
7
7
7
7
7
7
7
7
7
7
8
8
8
8
8

1
1
2
2
3
3
3
3
4
4
5
5
5
5
5
6
6
7
7
0.
0.
0.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.

0.
0.
0.
0.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
1.
2.
2.
2.
2.
,930
,930
,930
,326
,326
,326
,326
,326
,326
,326
,326
,326
,326
,786
,786
,786
,786
,786
,786
,786
,786
,786
,786
,786
,786
,786
,786
,786
,786
,312
,312
,312
,312
,312
,312
,312
,312
,312
,312
,312
,312
,312
,902
,902
,902
,902
,902

,494
,494
,859
,859
,150
,150
,150
,150
,425
,425
,732
,732
,732
,732
,732
,099
,099
,543
,543
1
0
1
0
1
3
1
0
1
1
0
0
1
3
6
1
0
0
1
2
0
2
3
5
0
3
3
3
5
0
1
2
3
2
3
5
0
2
5
1
2
1
1
4
3
8
2

0
0
0
2
3
1
2
1
3
0
0
4
0
1
0
3
2
3
2
0.0650
-0.8699
0.0650
-1.0004
-0.2458
1.2632
-0.2458
-1.0004
-0.2458
-0.2458
-1.0004
-1.0004
-0.2458
0.7656
2.6578
-0.4959
-1.1267
-1.1267
-0.4959
0.1348
-1.1267
0.1348
0.7656
2.0271
-1.1267
0.7656
0.7656
0.7656
2.0271
-1.2513
-0.7100
-0.1688
0.3725
-0.1688
0.3725
1.4551
-1.2513
-0.1688
1.4551
-0.7100
-0.1688
-0.7100
-0.9020
0.5204
0.0463
2.4170
-0.4279

-0.9887
-0.9887
-1.0732
1.4251
1.7287
-0.1400
0.7944
-0.1400
1.1858
-1.0729
-1.0898
1.4275
-1.0898
-0.4604
-1.0898
0.4839
-0.0534
0.2128
-0.2530
C-35
DRAFT - DO NOT CITE OR QUOTE

-------
1
2
3
4
5
6
7
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
134.
134.
134.
134.
134.
134.
134.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.

2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000

,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
7.
7.
7.
7.
8.
8.
8.
2.
3.
4.
4.
5.
5.
5.
5.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
7.
7.
7.
7.
7.
7.
7.
7.
7.
7.
7.
7.
8.
8.
8.
9.
9.

1.
1.
1.
2.
2.
3.
3.
3.
3.
4.
4.
4.
4.
4.
4.
4.
5.
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000

0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
.363
.363
.363
.363
.383
.383
.383
.703
.631
.562
.562
.506
.506
.506
.506
.466
.466
.466
.466
.466
.466
.466
.466
.466
.466
.466
.466
.466
.466
.466
.466
.444
.444
.444
.444
.444
.444
.444
.444
.444
.444
.444
.444
.437
.437
.437
.443
.443

.926
.926
.926
.894
.894
.853
.853
.853
.853
.802
.802
.802
.802
.802
.802
.802
.743
7
7
7
7
8
8
8
2
3
4
4
5
5
5
5
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
7
7
7
7
7
7
7
7
7
7
7
7
8
8
8
9
9

1
1
1
2
2
3
3
3
3
4
4
4
4
4
4
4
5
2.
2.
2.
2.
3.
3.
3.
1.
1.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
2.
3.
3.
3.
3.
3.
3.
3.
3.
3.
3.
3.
3.
3.
3.
3.
3.
3.

0.
0.
0.
1.
1.
2.
2.
2.
2.
3.
3.
3.
3.
3.
3.
3.
3.
,543
,543
,543
,543
,068
,068
,068
,406
,892
,250
,250
,530
,530
,530
,530
,796
,796
,796
,796
,796
,796
,796
,796
,796
,796
,796
,796
,796
,796
,796
,796
,105
,105
,105
,105
,105
,105
,105
,105
,105
,105
,105
,105
,496
,496
,496
,985
,985

,926
,926
,926
,789
,789
,559
,559
,559
,559
,208
,208
,208
,208
,208
,208
,208
,714
2
2
2
0
2
0
8
2
3
2
1
3
5
3
1
3
3
3
5
6
5
2
0
2
0
5
4
3
2
4
2
0
4
5
1
4
1
5
3
4
1
3
3
0
7
5
0
6

1
1
1
1
2
3
1
1
3
4
4
4
2
3
4
4
1
-0.2530
-0.2530
-0.2530
-1.1847
-0.4373
-1.2562
2.0195
0.8346
1.1101
-0.1967
-0.9842
0.3091
1.6241
0.3091
-1.0058
0.1162
0.1162
0.1162
1.2556
1.8253
1.2556
-0.4534
-1.5928
-0.4534
-1.5928
1.2556
0.6859
0.1162
-0.4534
0.6859
-0.4534
-1.5658
0.4511
0.9554
-1.0615
0.4511
-1.0615
0.9554
-0.0531
0.4511
-1.0615
-0.0531
-0.0531
-1.5793
1.5832
0.6796
-1.6270
0.8225

0.2834
0.2834
0.2834
-1.5502
0.4157
0.5454
-1.9294
-1.9294
0.5454
0.6851
0.6851
0.6851
-1.0440
-0.1795
0.6851
0.6851
-1.7660
C-36
DRAFT - DO NOT CITE OR QUOTE

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005
2005

.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000

5.
5.
5.
5.
5.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
6.
7.
7.
7.
7.
7.
8.

Combine litters
within dose
the

fit of

0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000
0000

0.743
0.743
0.743
0.743
0.743
0.681
0. 681
0. 681
0.681
0.681
0. 681
0. 681
0.681
0.681
0. 681
0. 681
0.681
0.623
0. 623
0. 623
0. 623
0.623
0.576

























with adjacent levels
groups until
the

the expected
XA2 statistic to


5
5
5
5
5
6
6
6
6
6
6
6
6
6
6
6
6
7
7
7
7
7
8

of the
count
3.714
3.714
3.714
3.714
3.714
4.086
4.086
4.086
4.086
4.086
4.086
4.086
4.086
4.086
4.086
4.086
4.086
4.361
4.361
4.361
4.361
4.361
4.606

3
5
5
4
4
6
2
4
5
6
5
4
5
3
6
0
0
7
5
5
7
6
0

-0
0
0
0
0
0
-1
-0
0
0
0
-0
0
-0
0
-2
-2
1
0
0
1
0
-1

.4648
.8364
.8364
.1858
.1858
.9945
.0836
.0445
.4750
.9945
.4750
.0445
.4750
.5641
.9945
.1227
.1227
.1486
.2781
.2781
.1486
.7133
.7419

litter-specific covariate
exceeds 3
.0, to help
improve
chi- square.






Grouped Data




0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

134
134
134
134
134
134
134
134

Dose


.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000

.0000
.0000
.0000
.0000
.0000
.0000
.0000
.0000

Lit





































Mean

.-Spec. Cov.


1.0000
2.0000
3.0000
4.0000
5.0000
5.0000
5.0000
5.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
8.0000
8.0000
8.0000

1.0000
2.0000
3.0000
3.0000
4.0000
5.0000
5.0000
5.0000








































Expected Observed


0.
1.
2.
3.
3.
3.
3.
1.
3.
3.
3.
3.
3.
3.
3.
3.
4.
4.
4.
4.
4.
4.
2.
5.
5.
2.

0.
1.
3.
1.
2.
3.
3.
1.


270
675
401
722
977
977
977
326
573
573
573
573
573
573
573
573
624
624
624
624
624
624
312
805
805
902

989
718
449
150
850
463
463
732


0
3
2
2
4
2
1
1
9
1
1
2
5
5
6
8
1
5
5
5
7
3
1
5
11
2

0
2
6
1
3
4
1
0
Scaled
Residual


-0.5586
1.0237
-0.2443
-0.8049
0.0098
-0.8614
-1.2970
-0.2458
2.4207
-1.1474
-1.1474
-0.7013
0.6367
0.6367
1.0827
1.9747
-1.3869
0.1441
0.1441
0.1441
0.9096
-0.6214
-0.7100
-0.2698
1.7418
-0.4279

-1.3982
0.2488
1.3759
-0.1400
0.0799
0.2388
-1.0962
-1.0898












































































C-37
DRAFT - DO NOT CITE OR QUOTE

-------
1
2
3
4
5
6
7
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
134.
134.
134.
134.
134.
134.
134.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.
526.

2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
2005.
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000

,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
,0000
6.0000
7.0000
7.0000
7.0000
8.0000
8.0000
8.0000
2.0000
3.0000
4.0000
5.0000
5.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
7.0000
8.0000
8.0000
8.0000
9.0000
9.0000

1.0000
2.0000
3.0000
3.0000
4.0000
4.0000
4.0000
4.0000
4.0000
4.0000
4.0000
5.0000
5.0000
5.0000
5.0000
5.0000
5.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
6.0000
4.199
5.086
5.086
5.086
3.068
3.068
3.068
1.406
1.892
4.500
5.060
5.060
5.592
5.592
5.592
5.592
5.592
5.592
5.592
5.592
3.105
3.105
3.105
3.105
3.105
3.105
3.105
3.105
3.105
3.105
3.105
3.105
3.496
3.496
3.496
3.985
3.985
2.777
3.577
5.118
5.118
3.208
3.208
3.208
3.208
3.208
3.208
3.208
3.714
3.714
3.714
3.714
3.714
3.714
4.086
4.086
4.086
4.086
4.086
4.086
4.086
4.086
4.086
4.086
4.086
5
5
4
2
2
0
8
2
3
3
8
4
6
8
11
2
2
9
5
6
0
4
5
1
4
1
5
3
4
1
3
3
0
7
5
0
6
3
3
4
4
4
4
4
2
3
4
4
1
3
5
5
4
4
6
2
4
5
6
5
4
5
3
6
0
0
-0
-0
-1
-0
-1
2
0
1
-0
1
-0
0
0
2
-1
-1
1
-0
0
-1
0
0
-1
0
-1
0
-0
0
-1
-0
-0
-1
1
0
-1
0
0
-0
-0
-0
0
0
0
-1
-0
0
0
-1
-0
0
0
0
0
0
-1
-0
0
0
0
-0
0
-0
0
-2
.3044
.0284
.3578
.0166
.4373
.2562
.0195
.8346
.1101
.8351
.3670
.4926
.1644
.9700
.1785
.4469
.4469
.3729
.2384
.1644
.5658
.4511
.9554
.0615
.4511
.0615
.9554
.0531
.4511
.0615
.0531
.0531
.5793
.5832
.6796
.6270
.8225
.4909
.8022
.9786
.9786
.6851
.6851
.6851
.0440
.1795
.6851
.6851
.7660
.4648
.8364
.8364
.1858
.1858
.9945
.0836
.0445
.4750
.9945
.4750
.0445
.4750
.5641
.9945
.1227
C-38
DRAFT - DO NOT CITE OR QUOTE

-------
1 2005.0000 6.0000 4.086 0 -2.1227
2 2005.0000 7.0000 4.361 7 1.1486
3 2005.0000 7.0000 4.361 5 0.2781
4 2005.0000 7.0000 4.361 5 0.2781
5 2005.0000 7.0000 4.361 7 1.1486
6 2005.0000 7.0000 4.361 6 0.7133
7 2005.0000 8.0000 4.606 0 -1.7419
9 Chi-square = 105.13 DF = 98 P-value = 0.2930
10
11 To calculate the BMD and BMDL, the litter specific covariate is fixed
12 at the mean litter specific covariate of all the data: 5.379518
13
i -j 	
Specified effect = 0.05
Risk Type = Extra risk
Confidence level = 0.95
BMD = 67.0227
BMDL = 44.6514
0.8
0.7
1
g 0.5
t3
ro
0.4
0.3
0.2
Nested Logistic Model with 0.95 Confidence Level
	 Nr'-trri I nni~tir 	
T -
^^^-—^
T /K^^^ ^
"I'
BMDL BMD
0 500 1000 1500 2000
dose
11:42 11/162007
          Figure C-6. Nested Logistic Model, 0.05 Extra Risk - Incidence of Cervical Rib in
          Mice versus Cmax Methanol, GD 6-15 inhalational study.
            Source:  Rogers et al. (1993).

14          Once the BMDLos was obtained in units of mg/L, it was used to derive a chronic
15   inhalation reference value.  The first step is  to calculate the HEC using the PBPK model
16   described in Appendix B. An algebraic equation is provided (Equation 1 of Appendix B) that
17   describes the relationship between predicted methanol AUC and the human equivalent inhalation
18   exposure concentration (HEC) in ppm. This equation can also be used to estimate model
19   predictions for HECs from  Cmax values because Cmax values and AUC values were estimated at
20   steady-state for constant 24-hour exposures (i.e., AUC = 24 x Cmax).

21         BMDLHEC (ppm) = 0.0224*BMDL05*24+(1334*BMDLo5*24)/(794+ BMDL05*24)

                                              C-39          DRAFT - DO NOT CITE OR QUOTE

-------
 1         BMDLHEC (ppm) = 0.0224*44.7*24 + ((1334*44.7*24)7(794+ 44.7*24)) = 791 ppm

 2          Next, because RfCs are typically expressed in units of mg/m3, the HEC value in ppm was
 3    converted using the conversion factor specific to methanol of 1 ppm =1.31 mg/m3:

 4                          HEC (mg/m3) = 1.31 x 791 ppm = 1036 mg/m3

 5          Finally, this HEC  value was divided by a composite 100-fold UF (3 for uncertainty
 6    associated with animal to human differences, 10 for consideration of human variability, and 3 for
 7    database deficiencies) to obtain the chronic inhalation reference value:

 8                         RfC (mg/m3) = 1036 mg/m3 - 100 =  10.4 mg/m3

      C.4. RFC DERIVATIONS USING BURBACHER ET AL. (1999A,B)
 9          The BMD approach was utilized in the derivation of potential chronic inhalation
10    reference values from effects seen in monkeys due to prenatal methanol exposure.  Deficits in
11    VDR were evaluated from Burbacher et al. (1999a, 1999b). In the application of the BMD
12    approach, continuous models in the EPAs BMDS, version 1.4. Ic, were fit to the dataset for
13    increased latency in VDR in neonatal monkeys. As the EPAs PBPK model was not
14    parameterized for monkeys, external concentration (ppm) was used as the dose metric.
15          The VDR test, which assesses time (from birth) it takes for an infant to grasp for a
16    brightly colored object containing an applesauce-covered nipple, is a measure of sensorimotor
17    development. Beginning at 2 weeks after birth, infants were tested 5 times/day, 4 days/week.
18    Performance on that test,  measured as age from birth at achievement of test criterion (successful
19    object retrieval on  8/10 consecutive trials over 2 testing sessions), was reduced in all treated male
20    infants. The times (days after birth) to achieve the criteria for the VDR test were 23.7 ±4.8
21    (n = 3), 32.4 ± 4.1  (n = 5), 42.7 ± 8.0 (n = 3), and 40.5 ± 12.5 (n = 2) days for males and
22    34.2 ± 1.8 (n = 5),  33.0 ± 2.9 (n = 4), 27.6 ± 2.7 (n = 5), and 40.0 ± 4.0 (n = 7) days for females
23    in the control to 1800 ppm groups, respectively. As discussed in Section 4.3.2, this type of
24    response data is sometimes adjusted to account for premature births by subtracting time (days)
25    premature from the time (days from birth) needed to meet the test criteria (Wilson and Cradock,
26    2004). When this type of adjustment is applied, the times (days after birth or, if shorter, days
27    after control mean  gestation length) to achieve the criteria for VDR test were 22.0 ± 9.54  (n = 3),
28    26.2 ± 8.61 (n = 5), 33.3 ± 10.0 (n = 3), and 39.5 ± 16.3 (n = 2) days for males and 32.0 ± 4.3
29    (n = 5), 21.8 ± 5.6  (n = 4), 24.0 ± 5.7 (n = 5), and 32.0 ± 14.8 (n = 7) days for females in the
30    control to 1800 ppm groups, respectively.  When these data were modeled within BMDS
31    (version 2.1), there was no significant difference between unadjusted responses and/or variances
32    among the dose levels for males and females combined (p = 0.244), for males only (p = 0.321)
                                               C-40           DRAFT - DO NOT CITE OR QUOTE

-------
 1    and for males only with the high-dose group excluded (p = 0.182), or for adjusted responses of
 2    males and females combined (p = 0.12), males only (p = 0.448) and males only with the high-
 3    dose group excluded (p = 0.586).91 The only data that offered a significant dose-response trend
 4    was that for unadjusted (p = 0.0265) and adjusted (p = 0.009) female responses, but the model
 5    fits for the adjusted female response data were unacceptable.  Only the unadjusted female VDR
 6    response data offered both a dose-response trend and acceptable model fits.  The modeling
 7    results for this data set are presented in Table C-10.
 8          The current BMD technical guidance (U.S. EPA, 2000b) suggests that in the absence of
 9    knowledge as to what level of response to  consider adverse, a change in the mean equal to
10    1 control S.D. from the control mean can be used as a BMR for continuous endpoints. A
11    summary of the results most relevant to the development of a POD using the BMD approach
12    (BMD, BMDL, and model fit statistics) for increased latency of VDR in female neonatal
13    monkeys exposed to methanol with a BMR of 1 control mean S.D. is provided in Table C-10.
14    Model fit was determined by statistics (AIC and $ residuals of individual dose groups) and
15    visual inspection, as recommended by EPA (2000b).  The 3rd degree polynomial model returned
16    a lower AIC than the other models.92 The  BMDLiso was determined to be 81.7 hrxmg/L, using
17    the 95% lower confidence limit of the dose-response curve expressed in terms of the ppm of
18    external methanol concentration.
      91 BMDS continuous models contain a test for dose-response trend, test 1, which compares a model that fits a
      distinct mean and variance for each dose group to a model that contains a single mean and variance. The dose
      response is considered to be significant if this comparison returns ap value < 0.05.
      92 A detailed analysis of this dose response revealed that modeling results, particularly the BMDL estimation, are
      very sensitive to the high-dose response. There is no data to inform the shape of the curve between the mid- and
      high-exposure levels, making the derivation of a BMDL very uncertain.  The data were analyzed without the high
      dose to determine if the downward trend in the low- and mid-exposure groups is significant. It was not, so
      nonnegative restriction on the (3 coefficients of the poly models was retained.
                                                 C-41             DRAFT - DO NOT CITE OR QUOTE

-------
     Table C-10. Comparison of BMD modeling results for VDR in female monkeys using
     AUC blood methanol as the dose metric
Model
Linear
2nd degree polynomial
3rd degree polynomial
Powerb
Hill
BMD1SD (AUC,
hr x mg/L)a
119.058
114.094
120.176
133.517
132.283
BMDL1SD
(AUC,
hr x mg/L)a
51.9876
59.6412
81.6513
63.0615
-
p-value
0.1440
0.2388
0.2718
0.1112
NA
AICC
110.4492
109.43782
109.17894
111.11010
113.11010
Scaled residual11
0.5380
0.0994
0.0199
0.0000
0.0000
 1
 2
 O
 4
 5
 6
 7
 8
 9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
aAUC was estimated using a rat PBPK model as described in section 3.4 of the methanol lexicological review; the
BMDL is the 95% lower confidence limit on the AUC of a decrease of 1 control mean S.D. estimated by the model
using the likelihood profile method (U.S. EPA, 2000b).
bModel choice based on adequate p value (> 0.1), visual inspection, low AIC, and low (absolute) scaled residual.
°AIC = Akaike Information Criterion = -2L + 2P, where L is the log-likelihood at the maximum likelihood
estimates for the parameters, and P is the number of modeled degrees of freedom (usually the number of parameters
estimated).
dj^d residual (measure of how model-predicted responses deviate from the actual data) for the dose group closest to
the BMD scaled by an estimate of its S.D. Provides a comparative measure of model fit near the BMD. Residuals
that exceed 2.0 in absolute value should cause one to question model fit in this region.

Source:  Burbacheretal. (1999a).
 Polynomial  Model.
 (Version:  2.13;  Date:  04/08/2008)
 Input  Data  File: C:\USEPA\BMDS2\Data\Burbacher\PolfemSet.(d)
 Gnuplot  Plotting File:   C:\USEPA\BMDS2\Data\Burbacher\PolfemSet.plt
                                                 Fri  Dec 12 15:30:29  2008


 VDR in female monkeys using AUC blood methanol  as the dose metric


    The form of the response function is:


   Y[dose]  = beta_0 + beta_l*dose + beta_2*dose/"2 + ...


   Dependent variable =  F_VDR
   Independent variable  = F_Dose
   The  polynomial coefficients are restricted to be positive
   The  variance is to be modeled as Var(i)  =  exp(lalpha + log(mean(i)) * rho)


   Total  number of dose  groups = 4
   Total  number of records with missing values = 0
   Maximum number of iterations = 250
   Relative  Function Convergence has been  set to: le-008
   Parameter Convergence has been set to:  le-008


                   Default Initial Parameter Values
                           lalpha =      4.07254
                              rho =             0
                           beta_0 =          34.2
                           beta_l =             0
                           beta_2 =             0
                           beta_3 =             0


           Asymptotic Correlation Matrix of Parameter Estimates


            ( *** The model parameter(s)  -beta_l    -beta_2
                                            C-42
                                                                  DRAFT - DO NOT CITE OR QUOTE

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
JT
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
have been estimated at a boundary po
the user,
and do not appear in the correlation

lalpha rho beta 0

lalpha 1 -1 -0.0076
rho -1 1 0.0076
beta 0 -0.0076 0.0076 1
beta_3 0.018 -0.018 -0.37

Parameter Estimates

Variable Estimate Std. Err.
Limit
lalpha -13.5062 9.81148
5.72395
rho 4.90831 2.77841
10.3539
beta 0 31.5013 1.49057
34.4228
beta 1 8.36431e-025 NA
beta 2 0 NA
beta 3 3.19775e-006 1.53534e-006
006

NA - Indicates that this parameter has hit a bound
implied by some inequality constraint and thus
has no standard error.

Table of Data and Estimated Values of Interest

Dose N Obs Mean Est Mean Obs Std Dev


0 5 34.2 31.5 4.09
6.73 4 33 31.5 5.83
28.28 5 27.6 31.6 5.94
138.1 7 40 39.9 10.7

Model Descriptions for likelihoods calculated

Model Al: Yij = Mu(i) + e(ij)
Var{e(ij)} = Sigma/x2
Model A2 : Yij = Mu(i) + e(ij)
Var{e(ij)} = Sigma (i) ^2
Model A3: Yij = Mu(i) + e(ij)
Var{e(ij)} = exp(lalpha + rho*ln (Mu (i) ) )
Model A3 uses any fixed variance parameters that
were specified by the user

Model R: Yi = Mu + e(i)
Var{e (i) } = Sigma/x2

Likelihoods of Interest

Model Log (likelihood) # Param' s
Al -51.042924 5
A2 -47.867444 8
A3 -49.286738 6
fitted -50.589469 4
R -55.013527 2

Explanation of Tests

Test 1: Do responses and/or variances differ among
(A2 versus R)
int, or have been

matrix )

beta_3

0.018
-0.018
-0.37
1


95.0% Wald CI
Lower Conf. Limit

-32.7363

-0.537284

28.5798



1.88544e-007








Est Std Dev Sea


5.55
5.55
5.58
9.93 0

















AIC
112.085848
111.734888
110.573475
109.178938
114.027055



Dose levels?

specified by












Upper Conf.









6.20695e-








led Res.


1.09
0.54
-1.59
.0199




























C-43
DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
 Test 2:  Are Variances Homogeneous?  (Al vs A2)
 Test 3:  Are variances adequately modeled? (A2 versus A3)
 Test 4:  Does the Model for the Mean Fit?  (A3 versus fitted)
 (Note:  When rho=0 the results of Test 3 and Test 2 will be the same.)
   Test


   Test 1
   Test 2
   Test 3
   Test 4
                     Tests of Interest
-2*log(Likelihood Ratio)   Test df
            14.2922
            6.35096
            2.83859
            2.60546
  p-value


0.02654
0.09573
 0.2419
 0.2718
The p-value for Test 1 is less than .05.  There appears to be a
difference between response and/or variances among the dose levels
It seems appropriate to model the data


The p-value for Test 2 is less than .1.  A non-homogeneous variance
model appears to be appropriate


The p-value for Test 3 is greater than  .1.  The modeled variance appears
 to be appropriate here


The p-value for Test 4 is greater than  .1.  The model chosen seems
to adequately describe the data
      Specified  effect  =
      Risk  Type
      Confidence level  =
                   BMD  =
                  BMDL  =
                       Estimated S.D.s from the control mean
                            0.95
                          120.176
                          81.6513
                                      Polynomial Model with 0.95 Confidence Level
                      50
                      45
                      40
                      35
                      30
                      25
                      20
                                 Polynomial
                                                    BMDL
                                                                      BMD
                                  20
                                         40
                                                60
                                                       80
                                                              100
                                                                     120
                                                                            140
                                                  dose
                  15:30 12/122008
          Figure C-7. 3rd Degree Polynomial Model, BMR of 1 Control Mean S.D. - VDR in
          female monkeys using AUC blood methanol as the dose metric.
                                               C-44
                                                         DRAFT - DO NOT CITE OR QUOTE

-------
            Source: Burbacher et al. (1999a, 1999b)

 1          Once the BMDLiso was obtained in units of ppm, it was used to derive a chronic
 2   inhalation reference value. The first step is to calculate the HEC using the PBPK model
 3   described in Appendix B. An algebraic equation is provided (Equation 1 of Appendix B) that
 4   describes the relationship between predicted methanol AUC and the human equivalent inhalation
 5   exposure concentration (HEC) in ppm.

 6            BMDLHEC (ppm) = 0.0224*BMDLisD+(1334*BMDLisD)/(794+ BMDLiso)
 7              BMDLHEC (ppm) = 0.0224*81.7+(1334*81.7)7(794+ 81.7) = 126.3 ppm

 8          Next, because RfCs are  typically expressed in units of mg/m3, the HEC value in ppm was
 9   converted using the conversion factor specific to methanol of 1 ppm =1.31 mg/m3:

10                         HEC (mg/m3) = 1.31 x 126.3 ppm = 165 mg/m3

11          Finally, this HEC value  was  divided by a composite 100-fold UF  (3 for uncertainty
12   associated with animal to human differences, 10 for consideration of human variability, and 3 for
13   database deficiencies) to obtain the chronic inhalation reference value:

14                         RfC (mg/m3) = 165 mg/m3 -100 = 1.7 mg/m3
                                             C-45           DRAFT - DO NOT CITE OR QUOTE

-------
             APPENDIXD. RfC DERIVATION- COMPARISON OF DOSE METRICS

     D.I. METHODS
     D. 1.1.  Dose Metric Comparisons
 1          Three potential dose metrics were evaluated for possible use in risk extrapolation of
 2   methanol-induced developmental effects: AUC of methanol in the blood; Cmax of methanol in the
 3   blood; and total metabolism of methanol. The latter metric was considered because
 4   developmental effects may be caused by metabolites of methanol, particularly formaldehyde, and
 5   formate. These three metrics were evaluated by determining how well they were able to explain
 6   the variation in response for incidence of cervical ribs (CR) and supernumerary ribs (SNR) in a
 7   concentration-time bioassay by Rogers et al. (1995, raw data obtained from personal
 8   communication). In particular, pregnant CD-I mice were exposed to 2,000, 5,000, 10,000, or
 9   15,000 ppm methanol for 1, 2, 3, 5, or 7 hours on GD7 and developmental effects evaluated at
10   GDI 7.  This endpoint was selected because it was the most sensitive of those examined and gave
11   a reasonable dose-response relationship overall.
12          Initially, the fraction of pups within each litter carrying either or both CR and SNR was
13   calculated, and then the average across all litters in each concentration-time combination was
14   computed. However, as shown in Figure D-l, the resulting data appear to be nonmonotonic,
15   with the responses from 5-hour exposures exceeding those from 7-hour exposures, and the
16   responses from 2-hour exposures exceeding those from 3-hour exposures. It was noted that the
17   study was done with a block-design, where the dams/litters for some concentration-time
18   combination were divided between multiple blocks  and the average CR + SNR incidence in
19   controls varied from 30-52% among the 8 blocks. Therefore block-control response (percent)
20   was subtracted from each exposed litter's response (percent) before calculating an average
21   response among litters in a given concentration-time combination.  The resulting data are
22   presented in Figure D-1.
                                               D-1           DRAFT - DO NOT CITE OR QUOTE

-------
•\ n^ -T——^^

90

5? 75
OL
W 60
+
O 45

30

A: Raw data starting with ^
grand-mean of controls A- — ^^-'•^'-A
/^xrT o
/- ^ // , -
-^ /^^ r' / ~~ y<.
^ 9f f '
^ / ,''/'
, ./ f'f/
^^^v^^ z — ^


• 7-hr data
— A^ 5-hr data
— • — 3-hr data
0 2-hr data
X 1-hrdata
















15 -t 	 1 	 1 	 1 	 '
0 5000 10000 15000
Concentration (ppm)
75
60
5? 45
tt
W 30
+
0 15
0 C
-15
C

• 7-hr data
— A — 5-hr data
— • — 3-hr data
0 2-hr data
X 1-hrdata


/ *^ O
/X r//' -m
H-?^iK— —mf B: Data minus bkgd in
O each block, then summarized
) 5000 10000 15000
Concentration (ppm)




1
2
3
4
5
    Figure D-l. Exposure-response data for methanol-induced CR plus SNR
    malformations in mice at various concentration-time combinations.  The percent
       response in each litter was first calculated, with direct averages shown in the
       first panel relative to the grand-mean for the controls. In the second panel,
       the percent response in controls for each block of exposures in the study was
       first subtracted from each litter's response in that block before taking
       averages across litters.

       Source:  Rogers et al. (1995).

       While the correction for background differences does not completely correct the  apparent
nonmonotonic dose, the 2-hour response is now less than or below the 3-hour response at 5,000
and 10,000 ppm, and the strong disparity that appeared between the 5- and 7-hour data at 2,000
ppm is eliminated.  Overall, the data show a more consistent dependence on duration of
exposure, except for the response to 3 hours of 15,000 ppm methanol. Therefore these
                                         D-2
                                                            DRAFT - DO NOT CITE OR QUOTE

-------
 1
 2
 3
 4
 5
 6
 7
 8
 9
10
background-corrected response measures will be used to evaluate the 3 dose metrics, with the
exception that the 3-hour 15,000 ppm data point will be dropped as an outlier.  In particular, the
dose-response relationship based on these data will be plotted against each of the dose metrics to
determine which provides the most consistent overall dose-response relationship.

D.2. RESULTS
D.2.1. Dose Metric Comparisons
       The average incidence of CR plus SNR from the concentration-time developmental
bioassay of Rogers et al. (1995), with block-specific control values subtracted from each litter
average before calculating overall average responses, is plotted in Figure D-2 against three dose
metrics: AUC, Cmax, and total amount metabolized of methanol (The volume units for Cmax and
AUC were adjusted to put all three data sets on approximately the same scale for comparison).
ID
60
45
E
oTso

-------
 1   other two metrics. Thus, Cmax appears to be a better predictor of response than AUC or amount
 2   metabolized.  Looking at the exposure-response data in Figure D-l, one can see that 2- and
 3   3-hour exposures at 5,000 ppm elicit no increase over control, while 5- and 7-hour exposures at
 4   this level do.
 5          If AUC or amount metabolized were true measures of risk, then one would expect a
 6   graded response, where the 2- and 3-hour exposures were intermediate between controls and 5-
 7   7-hour exposures. But the lack of response at those shorter times  indicates that the concentration
 8   (Cmax) has not risen high enough in such a short exposure to cause a response, while it has at the
 9   longer durations. From Figure D-2, it appears that a Cmax of 11 mg/cL (1,100 mg/L) is a
10   NOAEL, with a linear increase in CR + SNR from that level to 38 mg/cL, after which the
11   response begins to plateau. Note that while the plot is of response above background, the plateau
12   is effectively at 100% total incidence: the highest points in Figure D-l are from the 7-hour
13   exposures at 15,000 ppm, where actual incidence was 98% (30% in controls); and the next
14   highest points are from the 5-hour 15,000 and 10,000 ppm exposures, where the incidences were
15   94% and 93%, respectively (32% in controls; both from the same  block).
                                               D-4           DRAFT - DO NOT CITE OR QUOTE

-------
          APPENDIX E. EVALUATION OF THE CANCER POTENCY OF METHANOL

      E.I. INTRODUCTION
 1          Two studies were selected for the evaluation of the cancer potency of methanol (Soffritti
 2    et al., 2002a, 2000b, 2002c; NEDO, 1987, 1987/2008b). The Soffritti et al. (2002a) study is the
 3    only oral study available with effects that show a statistically significant increase in incidence of
 4    any cancer endpoints in the treated groups versus the concurrent control group (pair-wise
 5    comparison) and is used to derive the POD for deriving an oral cancer slope factor. The NEDO
 6    (1987, 1987/2008b) 24-month rat study is the only inhalation study available with effects that
 7    show a statistically significant increase in incidence of any cancer endpoints and was used to
 8    derive the POD for the inhalation cancer unit risk. A third study, Apaja (1980), reported
 9    statistically significant increases in malignant lymphomas in Eppley Swiss Webster mice over
10    historical controls (pair-wise comparison) following drinking water exposure to methanol.
11    Because this study did not involve a concurrent control group it is not used for the derivation of a
12    cancer oral slope factor, but its dose-response is evaluated here for comparative purposes.

      E.I.I. Oral CSF POD
13          The Soffritti et al. (2002a) study, conducted by the Ramazzini Foundation, presents a
14    number of challenges if these data are to be used in dose-response modeling to assess the
15    carcinogenic potency  of methanol.  One challenge, determining the appropriate HED, is best
16    addressed using a PBPK model to derive an HED dose that considers the kinetic differences in
17    humans and the animal model, i.e., species extrapolation. Such a model was developed by the
18    EPA and is not addressed in this appendix; however, the dose metrics derived from that PBPK
19    model are used in the  modeling of the data.
20          The other major challenge, which is addressed in this appendix, is how to model the
21    nonstandard protocol by which methanol was tested, as reported in Soffritti et al.  (2002a). In
22    most oncogenicity studies, typified by those conducted by the NTP, animals are dosed for
23    104 weeks, with a scheduled sacrifice of all surviving animals at the end of treatment.  In the
24    study for methanol reported by Soffritti et al. (2002a), while the animals were treated with
25    methanol for 104 weeks, animals were not euthanized and examined on a specified schedule but
26    were followed until their natural death. It is well known that the incidence of background tumors
27    in a number of organs is different between that seen at a scheduled sacrifice at 104 or 105 weeks
28    and in the same sex/strain that is followed for a lifetime. A higher background incidence can
29    increase the difficulty of detecting chemically related responses (Melnick et al., 2007). Further,
30    performing pathological examinations on tissues collected after natural death can create
                                                E-1       DRAFT -DO NOT CITE OR QUOTE

-------
 1    difficulties associated with cell autolysis.93 At the same time, the shorter duration of the 2-year
 2    bioassays used at the NTP misses about two thirds of the life span of the rodent, potentially
 3    missing late stage or late appearing chemically related tumor responses (Melnick et al., 2007).
 4    ERF believes that "cutting short an experiment after two years may mask a possible carcinogenic
 5    response," but ERF further suggests that all chronic cancer studies "should continue until
 6    spontaneous animal death" (Soffritti et al., 2002a).  Soffritti et al. (2002a) cite ERF studies of
 7    benzene, xylenes, mancozeb, and vinyl acetate monomer as examples for which carcinogenic
 8    responses were observed after the 2-year treatment period.
 9          These data were evaluated using three different approaches and two different dose-
10    response models (EPA's multistage cancer and a multistage Weibull time-to-tumor model).
11    These approaches involved using the administered dose and models that would rely upon the
12    published data (Option 1), while another would rely on unpublished information that would be
13    provided by the Ramazzini Foundation (Options  1 and 2), while a third option (Option 3) would
14    also rely upon this unpublished information but would also incorporate results from PBPK
15    modeling for methanol developed by the EPA.

      E.I.2. Selection of the Data Modeled and HED
16          The individual animal data from the Ramazzini Foundation study was provided to EPA in
17    the standard NTP format in which the number of days  on study, the tissues examined and the
18    tumor types found were given for each animal. The tumors with incidences that were
19    statistically significantly increased or were considered to be rare tumors and considered for dose-
20    response modeling were the incidence of hepatocellular carcinoma in male rats and the incidence
21    of hemolymphoreticular neoplasms in both male  and female rats. The incidence of lympho-
22    immunoblastic lymphomas was modeled separately, and the combined incidence of all the
23    lymphomas was considered for dose-response modeling. Table E-2 provides the incidence of
24    these neoplasms reported in each dose group. The incidence of histiocytic sarcomas and myeloid
25    leukemias were not significantly increased in either sex.  The incidence of these tumors was not
26    combined with the lymphoblastic lymphomas because they are of a different  cell line and the
27    combination is not typically evaluated either for statistical significance or dose-response
28    modeling (McConnell et al.,  1986).
29          The drinking water concentrations  provided in  the Soffritti et al. (2002a) study were
30    converted to doses in mg/kg-day. Initially, an attempt was made to estimate the dose of methanol
31    to individual animals for development of an average dose; however, water consumption
32    information was available only on a cage-by-cage basis.  Based on the available information, the
33    average water consumption for each treatment group was calculated using  the available data
      93 Autolysis may develop in carcasses of animals if they are not processed immediately after death or if the animals
      become severely moribund prior to death.  These types of changes can compromise pathological diagnosis and
      subclassification of neoplasms.
                                                E-2            DRAFT - DO NOT CITE OR QUOTE

-------
 1   reported for weeks 1-104.  Although individual body weights were available, the corresponding
 2   intake was not available. The average body weight over the period of dosing for the experiment
 3   (using measurements taken on day 1-day 736) was calculated for each dosed group. A weighted
 4   average was calculated for the body weights using the number of animals for which body
 5   weights were recorded at each time point.  The average body weight and the average water
 6   consumption in (mL/day) were used to calculate the mg/kg-day doses. The equation used for
 7   this calculation is:

                   _^    ,   .,     ,   x   Dose(ppm)x WaterConsumption(mL/day)
                   Dose(mg / kg - day) =	rt^  '	—.
                                                1000 x Body Weight(kg)

 8          Table E-l provides the values used to obtain the mg/kg-day doses, as well as the resulting
 9   mg/kg-day doses. In addition, the average  and median times of death were calculated for each
10   group (both dosed and control), for the only the dosed groups combined (excluding the controls),
11   and for all the groups in the study combined (including control). These values were obtained
12   using the reported weeks on study for each animal.  One male rat (ID # 129) in the 20,000 ppm
13   group was not examined microscopically and was excluded from the time of death calculations
14   and all modeling. If this animal was included in the calculations for the average and median
15   times of death, the median time of death for all male rat dosed groups would increase from 97-
16   98 weeks; all of the other average and median times of death that include the 20,000 ppm group
17   do not change.
18          In the absence of a kinetic model, extrapolation from animal to human was based on the
19   default assumption of body weight374.  This extrapolation was applied to the animal POD
20   estimates to obtain the HEDs reported in Table E-l.  This extrapolation was calculated using the
21   average body weight of the dosed animals excluding controls (0.33 kg for the female rats and
22   0.51 kg for the male rats) over the dosing period of the study (through day 736) and 70 kg for the
23   human body weight. The equation used for the body weight4 extrapolation is
                                   [ Animal Body Weight (kg)
                                                           .1/4
                                   ^ Human Body Weight (kg) j
24    and results in a value of 0.26 for the female rats and 0.29 for the male rats.

       E. 1.2.1. Dose-Response Modeling
25    Option 1 - Quantal Dose-Response Modeling. Under this option, the standard default
26    modeling approach outlined in the Cancer Guidelines (U.S. EPA, 2005a) was applied. The
27    cancer bioassay data were fit using a multistage model, which is the current model preferred by
28    EPAs IRIS program for cancer dose-response modeling. The PODs for use in the derivation of
                                               E-3           DRAFT - DO NOT CITE OR QUOTE

-------
 1    CSFs were calculated using the multistage-cancer model available in the BMDS program
 2    (http ://www. epa. gov/NCEA/bmds) .
 3          BMDS was used to estimate BMDs and 95% lower bounds on the BMDs or BMDLs
 4    associated with a 10% extra risk (BMDLio). For this assessment, the multistage model was
 5    determinded to be an appropriate model for characterization of the dose-response curve in the
 6    observable range. At this time, the MOA for the tumors observed following exposure to
 7    methanol is not known; therefore, linear extrapolation was conducted to estimate a CSF.
 8          Option 2 - Time-to- Tumor Dose-Response Modeling. This option is similar to
 9    Option 1; however, rather than the use of quantal models, a time-to-tumor model was applied to
10    the selected datasets. Data for this analysis was provided by the Ramazzini Foundation and can
11    be obtained from their web site (http://www.ramazzini.it/fondazione/study.asp).  The same
12    assumptions regarding the FED and low-dose extrapolation were applied.  Because BMDS does
13    not include time-to-tumor modeling, the QRISK portion of Statox was relied upon for dose-
14    response modeling.  Statox is an internal EPA program that is used for gathering and analyzing
15    animal bioassay data and contains the QRISK component for dose-response modeling.  The
16    QRISK component of Statox Version 5.5 fits a multistage Weibull model to the data. The
17    multistage Weibull model is multistage in dose and Weibull in time and essentially assesses the
1 8    probability that a tumor would have been identified at time t. The multistage Weibull model has
19    the form:
                       p(d t} = 1 - e~v
20    with dose (d) and time (i) as the variables. The parameters estimated by fitting the model to the
21    data are the dose parameters go through q^ the induction time (to) and the power term for
22    time (c).
23          If t0 is interpreted as the time (assumed to be the same for all animals) from when a tumor
24    is observable (i.e., capable of being detected if the animal were to be sacrificed and a necropsy
25    performed) to the time the tumor causes the death of the animal, then these models can be
26    applied to data on incidental and fatal tumors simultaneously. Note that t and t0 only appear in
27    the model in the form of t-t0.  To make this explicit, we write P(d,t) = F(d,t-t0).  The probability
28    of an incidental tumor by time t is taken to be F(d,t) (t0 = 0) and the probability of a fatal tumor
29    by time t is taken to be F(d,t-t0). There are three possible types of incidence contexts for each
30    animal which contribute separately to the likelihood function for this model.  These are:
31       •  Censored response - animal died without having the tumor(s) being modeled
32       •  Incidental response - the animal died with the tumor(s) but the death  was not caused by
33          the tumor(s) (i.e., the time to death from those tumors would have been later than the
34          actual death time); and

                                              E-4           DRAFT - DO NOT CITE OR QUOTE

-------
 1       •  Fatal incidence - the tumor(s) being modeled was the cause of death.
 2          The contribution of each animal to the likelihood is then defined for its time of death ft).
 3    The complete likelihood is defined as:
       g
           Incidence(iJ)=Censored
Inc idenc e (i, j)=Inc idental
                                                                    n
Incidence(i,j)=Fatal
 4   where g is the number of dose groups in the study, including the control group, and /' varies from
 5    1 to the total number of animals in the study examined for the tumor type(s) being modeled.
 6          As with the quantal modeling, the lower bound on a dose at an extra risk of 10% was
 7   estimated. Goodness-of-fit was determined by visually inspecting graphical output of the
 8   modeling. AIC values were also calculated for the time-to-tumor model fit.
 9          This option was proposed because time-to-tumor modeling is typically applied to account
10   for differences in survival among treated and control groups. However, in this case there were
11   no differences detected in the survival times. Figures E-l  and E-2 are graphs  of the proportion
12   surviving versus the weeks on study for the female rats and male rats, respectively. In addition,
13   the Life Table program (Thomas et al.,  1997) was run on the data.  None of the statistical tests in
14   this program indicated a difference in survival between the control and the dosed groups.
15          The protocol used by the Ramazzini Foundation was different from that typically
16   employed in chronic rat bioassays. In typical rodent bioassays, a compound is administered to
17   the animals for  approximately 104 weeks, and the animals sacrificed within a short period (days)
18   following the end of treatment. In the Ramazzini Foundation bioassays (i.e., for methanol,
19   formaldehyde, MTBE, and aspartame), the animals were administered compounds for 104 weeks
20   but were allowed to live until a natural  death, which, in some animals, occurred months after the
21   completion of chemical administration.  This can have an impact on the tumor incidence and
22   therefore, the potential risk of tumor development associated with administration of a given
23   compound.  Time-to-tumor modeling was used in this case to attempt to adjust for the extended
24   life span of the  some of the animals in this study.
25          For time-to-tumor modeling, the POD must consider a specific time as well as a specific
26   risk level.  Since this study was not a standard study with a fixed study length, several
27   assumptions can be made with regard to the time to be used. For this modeling exercise, Two
28   different possible approaches were considered.
29          For the first approach, the model was fit to animal  data using all the times reported up to
30   the last death time or 153 weeks for the female rats and 148 weeks for the male rats. Every
31   tumor observed was assumed to be a fatal tumor or the cause of death in the animal. While the
32   animals lived longer than 104 or 105 weeks, the POD was calculated at 105 weeks, since it was
33   assumed that an animal life span of 148-153 weeks would not correspond to the average 70-year
34   human life span.

                                               E-5           DRAFT - DO NOT CITE OR QUOTE

-------
 1          For the second approach, an attempt was made to simulate what might have occurred if
 2   the study had been a standard 2-year protocol that was terminated at 105 weeks, with all
 3   surviving animals sacrificed at 105 weeks. It was assumed that the tumors discovered in the
 4   animals that survived longer than 105 weeks would have been present and found at necropsy.
 5   Therefore, all tumors in animals that died in weeks 105 and earlier were assumed to be fatal or
 6   the cause of death in the animals, and all tumors that would have been discovered at the necropsy
 7   of animals were assumed to be incidental or not the cause of death. The life span assumed for
 8   this analysis was 105 weeks in the rat and the POD was calculated for a 10% risk to a human at
 9   105 weeks.  This approach was conducted mainly for comparative purposes to evaluate the
10   potential impact on the POD if the study duration was shortened and for a more direct
11   comparison to the quantal PODs and serves only  as bounding exercise for the risk.
12          Option 3 - Dose-Response Modeling using PK Dose Metrics.  For the third approach,
13   PK dose metrics obtained from the PBPK model (Section 3.4) were used as the doses. Both
14   time-to-tumor modeling and quantal modeling was done. The Statox program was used to
15   estimate MLEs and lower bounds on dose associated with a 10% extra risk (LEDlOs), and
16   BMDS multistage model was used to estimate BMDs and 95% lower bounds on the BMDs or
17   BMDLs associated with a 10% extra risk (BMDL10). Each of the dose metrics, provided in
18   Table E-5, were used in this option of the dose-response modeling.

      E. 1.2.1. Results for Oral Slope Factor POD
19          Quantal Dose-Response Modeling (Option 1). Quantal dose-response modeling was
20   conducted using estimated mg/kg-day doses and the incidence of lympho-immunoblastic
21   neoplasms and the combined lymphomas for the female rat and the hepatocellular carcinomas,
22   the lympho-immunoblastic  neoplasms, and the combined lymphomas for the male rat. The
23   results of the quantal modeling for this option are given in Table E-3.  The multistage model gave
24   an adequate  fit (p value > 0.05) and was able to derive BMDLio values for all of the lymphoma
25   data.  However, for the male rat hepatocellular carcinomas, the multistage model failed to
26   estimate a BMDLio. Human equivalent BMDLio values94  and CSF are also provided in
27   Table E-3.
28          The POD values calculated from the female rat data ranged from 259-405 mg/kg-day for
29   the lympho-immunoblastic  neoplasms and 251-400 mg/kg-day for the combined lymphomas.
30   For the HED POD values calculated from the male rat data, the range of values for the liver
31   hepatocellular carcinomas was 537-567 mg/kg-day; however, for 3 of the models (Multistage,
32   Quantal Linear and Log-Probit), the BMDS program failed to estimate a BMDLio. The lympho-
33   immunoblastic endpoint in the male rats produced POD values that ranged from 110-223 mg/kg-
34   day, and the all lymphomas incidence data gave POD values of 113-227 mg/kg-day.  Figures E-3
     94 Computed from the animal values by multiplying by the body weight'7' animal-to-human extrapolation value (0.26
     for females and 0.29 for males)
                                              E-6           DRAFT - DO NOT CITE OR QUOTE

-------
 1   through E-7 show graphs of the quantal models that give the lowest POD for each of the
 2   endpoints modeled.
 3          Time-to-Tumor Dose-Response Modeling (Option 2). Results of the time-to-tumor
 4   modeling using estimated mg/kg-day doses with POD and CSF values for each approach
 5   described above are given in Table E-4. Approach 2 gives smaller POD estimates than
 6   Approach 1. With Approach 2, an artificial end of the study is assumed of 105 weeks and the
 7   designation of approximately half of the total tumors changed from fatal to incidental. This
 8   approach evaluates the potential impact on the POD of terminating the  study at 105 weeks, rather
 9   than allowing the animals to live until their natural death, assuming that the same animals
10   bearing tumors would be  "observed" at 105 weeks, rather than at later time points.  For
11   Approach 1, the model was fit to the actual observed weeks-on-study, and a time of 105 weeks
12   was used in calculating the POD.  For the female rat, the PODs based on the lympho-
13   immunoblastic neoplasms were 349 mg/kg-day for approach 1 and 179 mg/kg-day
14   forApproach 2.  The PODs for the combined lymphomas ranged were 321 and 198 mg/kg-day
15   for the 2 appraoches. For the PODs calculated from the male rat data, the values were 783 and
16   612 mg/kg-day for the hepatocellular carcinomas, 174 and 91 mg/kg-day for the lympho-
17   immunoblastic neoplasms, and 192 and 92 mg/kg-day for the combined lymphomas.  Figures E-
18   3 through E-7 show the modeling results for the time-to-tumor modeling using Approache  1
19   where the time is fixed at 105 weeks and the doses are allowed to vary. Figures E-8 and E-9
20   show Kaplan-Meier curves versus the model fit to the combined lymphoma data for the females
21   and males, respectively.  In these graphs each line corresponds to a specific dose, and time is
22   allowed to vary up to the  study end of 153 or 148 weeks. For the male  rat combined lymphomas
23   (Figure E-9), the multistage Weibull predicted values more closely match the Kaplan-Meier at
24   105 weeks than at the average life span (94 weeks) or the end of study (148 weeks). For the
25   female rat combined lymphomas, the closest match of the Kaplan-Meier curves to the model
26   predicted values appears to be around the average life span of 96 weeks.
27          The AIC values for the time-to-tumor modeling are all higher than those for the quantal
28   modeling. However, since the LLFs on which the AIC are based are different for time-to-tumor
29   versus quantal models, the AIC values from time-to-tumor models cannot be compared to those
30   from quantal models.
31          Dose-Response Modeling using PBPK Dose Metrics (Option 3). Both time-to-tumor
32   and quantal dose-response modeling were conducted using the incidence of lympho-
33   immunoblastic neoplasms and the combined lymphomas for the female rat and the hepatocellular
34   carcinomas, the lympho-immunoblastic neoplasms, and the combined lymphomas for the male
35   rat and the three PBPK dose metrics (blood methanol AUC, peak blood concentration, and the
36   total metabolized per day) obtained from the EPA's PBPK model (Table E-5).
37          Only Approach 1 was used for the time-to-tumor dose-response modeling using PBPK
38   dose metrics, and results are given in Table E-6a; HEDs are provided in Table E-6b.
                                              E-7           DRAFT - DO NOT CITE OR QUOTE

-------
 1          The results of the multistage quantal modeling are given in Table E-7. Most model runs
 2   gave an adequate fit to the data by the ^ Goodness of Fit p-value (e.g., p-values > 0.05),
 3   although for male rat hepatocellular carcinomas, the calculations using the AUC or amount
 4   metabolized dose metric were unable to converge for the model fit or the derivation of a BMDL.
 5   A plot of the model fit for male rat combined lymphomas using total methanol metabolized per
 6   day as the dose metric (the endpoint and dose metric used in the derivation of the oral CSF) is
 7   shown in Figure E-10; HEDs are provided in Table E-8 where the BMDL was calculated.

     E.2. INHALATION ORAL CANCER UNIT RISK POD
 8          The NEDO (1987, 1987/2008b) study was conducted using a standard protocol with
 9   exposure for 104 weeks, followed by sacrifice of all animals surviving to 104 weeks.  As with
10   the Soffretti (2002a) study, pharmacokinetic dose metrics for use in the dose-response
11   assessment were determined for the inhalation exposures using the EPAs PBPK model.

     E.2.1. Selection of the Data Modeled and HED
12          The individual  animal data from the NEDO (1987/2008b)  study were provided in a 2008
13   translation of the study from Japanese to English.  Although the translation provided the number
14   of days on study and the neoplastic responses seen in each animal, the translation did not provide
15   results if a tissue was examined histopathologically with no neoplastic responses.  This makes it
16   difficult to determine which of the individual animals were not examined, although the tables did
17   indicate that, for some of the animals, selected organs (specifically a few lungs in males and a
18   few adrenal glands in females) were not examined. Therefore, time-to-tumor analysis could not
19   be conducted with results from the inhalation data as was done with the oral data.  However,
20   survival analysis of all the data from the NEDO (1987/2008b) study did not indicate that there
21   were any survival problems (Figures E-ll and E-12). This suggests that a time-to-tumor analysis
22   is not necessary.
23          The tumors with significantly increased incidence that were considered for dose-response
24   modeling were the female rat adrenal gland pheochromocytomas and the male rat lung tumors
25   (papillary adenomas and adenocarcinomas  combined or papillary adenomas, adenocarcinomas
26   and adenomatosis combined); Table E-9 gives the  incidence of these tumors.

       E.2.1.1. Dose-Response Modeling
27          Quantal Dose-Response Modeling using Pharmacokinetic Dose Metrics. For the
28   selected endpoints from the NEDO (1987) study, only quantal dose-response modeling using
29   pharmacokinetic internal dose metrics estimated by the PBPK model (described in Section 3.4
30   and Appendix B) was conducted.  Each of the dose metrics provided in Table E-10 was used in
31   the BMDS software with the incidence data in Table E-9 to estimate the BMDs and 95% lower
32   confidence limits (BMDLs) associated with a 10% extra risk (BMDLio).
                                               E-8           DRAFT - DO NOT CITE OR QUOTE

-------
     E.2.2. Results for the IUR POD
 1          Quantal dose-response modeling was conducted using the incidence of adrenal gland
 2   phoechromocytomas in female rats and the combined incidence of lung adenomas and
 3   adenocarcinomas or lung adenomas, adenocarcinomas and adenomatosis in male rats using the
 4   pharmacokinetics dose metrics derived from the EPA's PBPK model. The results of this
 5   modeling are given in Table E-l 1. The multistage model gave an adequate fit to the data in all
 6   instances as determined by the £ goodness-of-fit/?-value (e.g., ^-values > 0.05).  A plot of the
 7   model fit for female rat pheochromocytomas using total methanol metabolized per day as the
 8   dose metric (the endpoint and dose metric used in the derivation of the IUR) is shown in
 9   Figure E-13. HECs are provided in Table E-12.

     E.3. ANALYSIS OF APAJA (1980) DRINKING WATER STUDY
10          The Apaja (1980) study was similar to the Soffritti et al.  (2002a) study in  that it was a life
11   span drinking water study. The primary differences are that Apaja (1980) used Eppley Swiss
12   Webster mice,  did not stop exposure at 104 weeks and did not employ an untreated concurrent
13   control group.  Methanol exposure groups of this study served as controls for malonaldehyde
14   exposed mice.  As with the Soffretti (2002a) study, pharmacokinetic dose metrics for use in the
15   dose-response  assessment were determined for the oral exposures using the EPA's PBPK mouse
16   model.

     E.3.1. Selection of the Data Modeled and HED
17          Individual animal data from the Apaja (1980) study were not available. Therefore, time-
18   to-tumor analysis could not be conducted. The tumors with significantly increased incidence that
19   were considered for dose-response modeling were malignant lymphomas in male and female
20   mice; Table E-13 gives the incidence of these tumors.

       E. 3.1.1. Dose-Response Modeling
21          Quantal Dose-Response Modeling using Pharmacokinetic Dose Metrics. A 1st degree
22   multisage model was used to evaluate the malignant lymphoma response from the Apaja (1980)
23   study versus pharmacokinetic dose metrics without background95 was conducted. Each of the
24   dose metrics provided in Table E-14 was used  in the BMDS software with the incidence data in
25   Table E-13 to estimate the BMDs  and 95% lower confidence limits (BMDLs) associated with a
26   10% extra risk (BMDLio).
     95 The assumption of zero background dose is consistent with what was done for the derivation of the oral CSF and
     EPA practice.
                                              E-9            DRAFT - DO NOT CITE OR QUOTE

-------
     E.3.2. Results for the IUR POD
 1          Quantal dose-response modeling was conducted using the incidence of malignant
 2   lymphoma in male and female Eppley Swiss Webster mice using the pharmacokinetics dose
 3   metrics derived from the EPA's PBPK model.  The results of this modeling are given in
 4   Table E-15.  The 1st degree multistage model gave an adequate fit to the data in all instances as
 5   determined by the $ goodness-of-fit/>-value (e.g., ^-values > 0.05). A plot of the model fits for
 6   male and female mice using total methanol metabolized per day as the dose metric is shown in
 7   Figure E-14. BMDLio HECs associated with each dose metric are provided in Table E-16.

     E.4. BACKGROUND DOSE ANALYSES
 8          The primary purpose of this cancer analysis is for the determination of cancer risk
 9   associated with increases in the levels of methanol or its metabolites (e.g., formate,
10   formaldehyde) over background. Thus, the PBPK model estimates of internal dose used in the
11   dose-response analyses described above do not describe or account for background levels of
12   methanol  or its metabolites. However, background levels of methanol may have contributed to
13   the response levels reported for some of the tumors associated with methanol.  If this
14   contribution is large, it could significantly impact cancer risk estimates.
15          Of the Soffritti et al. (2002a), NEDO (1985/2008b) and Apaja (1989) studies, only the
16   NEDO study reported methanol blood levels in test animals. Based on background levels of
17   methanol  in the blood of F344 rats reported by NEDO (1987; 1985/2008b) PBPK model
18   estimates  of AUC methanol (mg-h/L), Cmax methanol (mg/L) and total metabolites (mg/day) in
19   control animals were obtained. AUC methanol, Cmax methanol and methanol metabolite
20   background levels were estimated to be 97.16  mg-h/L, 4.05 mg/L and 7.79 mg/day for female
21   F344 rats  and 96.02, 4.00 and 10.81 mg/day for male F344 rats, respectively.
22          Available bioassays do not allow for the quantification of the relationship between
23   background levels and background responses,  but the "background dose" Multistage-cancer
24   model contained in version 2.1 of EPA's BMDS software can be used to estimate what
25   background dose would be necessary to explain the dose-response data evaluated in this
26   assessment if all of the background response were due to background levels of methanol or its
27   metabolites.  The results of dose-response modeling of cancer endpoints evaluated in this
28   assessment using the Multistage-cancer "background dose" model are shown in Table  E-17.
29          As can be seen from Table E-17, the Multistage-cancer-bgdose model predicts  that AUC
30   methanol, Cmax methanol  and methanol metabolite background doses necessary to explain the
31   dose-response data if all  of the background response were due to background levels  of methanol
32   or its metabolites are well above the actual background levels for these metrics  predicted by the
33   PBPK model. In the case of lymphoma responses in Sprague-Dawley rats, the background doses
34   for all metrics predicted by the Multistage-cancer-bgdose model are 20-2,000 fold higher than

                                              E-10          DRAFT - DO NOT CITE OR QUOTE

-------
1   the actual background doses predicted in F344 rats by the PBPK model. In the case of female
2   pheochromocytoma and male lung responses in F344 rats, the background doses predicted by the
3   Multistage-cancer-bgdose model are 3-5 fold higher than the actual background doses predicted
4   by the PBPK model. The fact that background dose estimates from the Multistage-cancer-bgdose
5   model are several-fold and sometimes more than 1,000-fold higher than the PBPK model
6   predictions suggests that the contribution of background levels of methanol or its metabolites to
7   the background responses of these tumors is relatively small.
                                            E-11           DRAFT - DO NOT CITE OR QUOTE

-------
           Table E-l.  Calculation of mg/kg-day doses

Dose (ppm)
0
500
5000
20000
Female Sprague-Dawley Rats
Body
weight
(kg)
0.33
0.33
0.33
0.34
Water
consump.
(g/day or
mL/day)
42.55
43.05
41.11
37.26
Dose
(mg/kg
-day)
0
66.0
624.1
2177
Average
time of
death
(wk)
98
96
94
98
Median
time of
death
(wk)
102
99
97
101
Male Sprague-Dawley Rats
Body
weight
(kg)
0.50
0.49
0.50
0.54
Water
consump.
(g/day or
mL/day)
52.57
52.06
52.58
48.32
Dose
(mg/kg -
day)
0
53.2
524
1780
Average
time of
death
(wk)
91
97
93
93
Median
time of
death
(wk)
91
98
93
100

Averaged
over all
dosed groups
(excluding
control)
Averaged
over all
groups
(including
control)
0.33
0.33




96
97
99
100
0.51
0.51




94
93
97
96
 Source: Soffretti et al. (2002a).
           Table E-2. Incidence for neoplasms considered for dose-response modeling
Dose
(ppm)
Dose
(mg/kg-
day)
Number of
animals
examined
Heptocellular
carcinomas
Histiocytic
sarcoma
Leukemia
monocytic
Leukemia
myloid
Lymphoma
lymphoblastic
Lymphoma
lympocytic
Lymphoma
lympho-
immunoblastic
All
lymphomas
combined
Female Sprague-Dawley rats
0
500
5000
20000
0
66.0
624.1
2177
100
100
100
100
Cochran Armitage Trend Test





1
2
2
3
0.19





3
3
3
3
0.5
0
1
1
1
0.3
0
1
0
0
0.8
9
17
19a
21a
0.04
9
19a
20a
22b
0.04
Male Sprague-Dawley rats
0
500
5000
20000
0
53.2
524
1780
100
100
100
99
Cochran Armitage Trend Test
0
2
2
3
0.10
2
4
1
1
0.9
1
0
0
0
0.8
8
4
6
1
0.98
1
3
1
1
0.7





16
24
28a
37b
0.0007
17
27
29a
38b
0.001
Tisher's Exact/>-value < 0.05; bFisher's Exact p-value < 0.01
Source: Soffretti et al. (2002a).
                                                 E-12
DRAFT - DO NOT CITE OR QUOTE

-------
     Table E-3. Results from multistage (1°) quantal modeling rat data using mg/kg-day
     exposures and default HED derivation method

Female
Sprague-
Dawley rat
Male
Sprague-
Dawley rat
All organs
lympho-
immunoblastic
All organs - all
lymphomas
Hepatocellular
carcinoma
All organs
lympho-
immunoblastic
All organs - all
lymphomas
AIC
359.16
371.95
72.84
455.67
468.79
p-value
0.19
0.11
0.38
0.34
0.24
Scaled
residual at
observed
dose closest
toBMD
-0.07
-0.07
Animal values
BMD10
2179.51
2141.88
BMDL10
1058.99
1033.69
Human values
Human
equivalent
BMDIV
277.5
270.9
CSF"
3.6E-4
3.7E-4
Failed8
0.14
0.11
714.26
744.35
448.43
456.11
131.0
133.3
7.6E-4
7.5E-4
aModel failed to optimize.
Calculated as CSF = O.I/Human BMDL10.

Source:  Soffretti et al. (2002a).
     Table E-4. Results from time-to-tumor modeling data using mg/kg-day exposures and
     default HED derivation method


AIC
Prediction
time (weeks)
Human values'"
MLE
(mg/kg-day)
LED10
(mg/kg-day)
CSFC
Approach 1 - Model fit to actual death times, dose estimates computed at 105 weeeks"
Female
Sprague-
Dawley rat
Male
Sprague-
Dawley rat
All organs lympho-immunoblastic
All organs all lymphomas
Hepatocellular carcinoma
All organs lympho-immunoblastic
All organs all lymphomas
831.42
900.46
105.45
1254.58
1309.86
105
105
105
105
105
729.0
679.1
4250.7
302.0
356.7
348.6
320.9
783.3
173.8
191.8
2.9E-4
3.1E-4
1.3E-04
5.8E-4
5.2E-4
Approach 2 - Truncating study at 105 Weeks3
Female
Sprague-
Dawley rat
Male
Sprague-
Dawley rat
All organs lympho-immunoblastic
All organs all lymphomas
Hepatocellular carcinoma
All organs lympho-immunoblastic
All organs - all lymphomas
631.81
664.43
1.16E+05d
914.42
935.29
105
105
105
105
105
370.7
431.9
1013.7
150.8
157.1
178.5
198.0
612.2
91.2
92.2
5.6E-4
5.1E-4
1.6E-4
1.1E-3
1.1E-3
Individual animal pathology data needed for the modeling reported in this table can be obtained from the Ramazzini Foundation web site
(http: //www. ramazzini. it/fondazione/study .asp).
bHuman values are computed by converting the animal doses to HED before modeling by multiplying by the body weight37' animal-to-
human extrapolation value (0.26 for females and 0.29 for males).
"Calculated as CSF = O.I/Human LED10.
dModel failed to optimize.

Source: Soffretti et al. (2002a).
                                               E-13
DRAFT - DO NOT CITE OR QUOTE

-------
    Table E-5. PBPK model estimated dose-metrics for doses
Sex
Female
Sprague-
Dawley
Male
Sprague-
Dawley
Dose (mg/kg-day)
66
624.1
2177
53.2
524
1780
Body weight
(kg)
0.33
0.33
0.34
0.49
0.50
0.54
AUC
(mg-h/L)
66.84
9543.63
91262.27
55.76
7500.14
80420.25
Peak (mg/L)
5.96
500.59
4157.98
4.81
395.56
3629.08
Amount metabolized
(mg/day)
18.39
126.68
141.60
21.82
168.85
200.03
Source: Soffretti et al. (2002a).
    Table E-6a. Results from time-to-tumor modeling of data using PBPK dose metrics
Approach 1 - Model fit
to actual death times,
dose estimates
computed at 105
weeeks
Female
Sprague-
Dawley
rat
Male
Sprague-
Dawley
rat
All organs
lympho-
immunoblastic
All organs all
lymphomas
Hepatocellular
carcinoma
All organs
lympho-
immunoblastic
All organs - all
lymphomas
Prediction
Time (wk)
105
105
105
105
105
AUC (mg-h/L)
AIC
832.43
901.43
110.47
1242.72
1297.74
MLE
152281
143138
undefined
55409
66211
LED10
64486
59574
119186
30059
33315
Peak (mg/L)
AIC
832.35
901.36
110.45
1244.61
1297.66
MLE
6763.4
6356.1
undefined
2460.2
2939.5
LED10
2895.3
2675.3
5385.9
1341.3
1487.2
Amount Metabolized
(mg/d)
AIC
829.24
898.24
109.60
1242.66
1295.95
MLE
160.45
145.53
undefined
141.63
143.14
LED10
90.7
82.2
300.4
77.3
82.7
 Source: Soffretti et al. (2002a)
    Table E-6b. HEDs from time-to-tumor modeling of data using PBPK dose metrics
Approach 1 - Model fit to actual death times,
dose estimates computed at 105 weeeks
Female
Sprague-
Dawley rat
Male
Sprague-
Dawley rat
All organs lympho-immunoblastic
All organs all lymphomas
Hepatocellular carcinoma
All organs lympho-immunoblastic
All organs - all lymphomas
AUC (mg-h/L)
LED10 (mg/kg-day)
647.0
618.3
965.4
594.0
657.1
Peak (mg/L)
LED10 (mg/kg-day)
676.2
645.4
1023.9
457.3
478.1
Amount
Metabolized (mg)
LED10 (mg/kg-day)
87.7
79.4
260.0
62.4
66.7
Source: Soffretti et al. (2002a)
                                        E-14
DRAFT - DO NOT CITE OR QUOTE

-------
               Table E-7.  Results of Multistage (1°) quantal modeling of data using PBPK dose
               metrics
AUC (mg-h/L)
AIC
P-
value
Scaled
residual
at dose
nearest
toBMD
BMD10
BMDL10
Peak (mg/L)
AIC
P-
value
Scaled
residual
at dose
nearest to
BMD
BMD10
BMDL10
Amount Metabolized (mg/d)
AIC
P-
value
Scaled
residual
at dose
nearest to
BMD
BMD10
BMDL10
Female Sprague-Dawley rats
All organs lympho-immunoblastic
357.9
0.13
-0.132
112790
49771
359.9
0.13
-1.56
5034.3
2240.3
357.9
0.34
-0.913
139.8
77.2
All organs - all lymphomas
372.78
0.075
-0.128
111621
48789.6
372.17
0.077
-0.145
4982.2
2196.3
370.57
0.18
-0. 002
134.4
74.5
Male Sprague-Dawley rats
Hepatocellular carcinoma
73.21
0.37
Failed8
73.19
0.37
Failed8
72.55
0.36
Failed8
All organs lympho-immunoblastic
457.45
0.14
1.214
36865
21989
457.3
0.15
1.173
1639.0
981.7
455.46
0.36
-0.750
94.4
61.6
All organs - all lymphomas
468.09
0.15
0.878
37822
22195
467.98
0.15
0.838
1686.5
993.1
467.35
0.20
-0.901
101.7
63.9
aBMD computation failed.  BMD is larger than three times maximum input doses.

Source: Soffretti et al. (2002a).
               Table E-8. Application of human PBPK model to derive HEDs from results of
               multistage (1°) quantal modeling of data using PBPK dose metrics

Female
Sprague-
Dawley rat
Male
Sprague-
Dawley rat

All organs lympho-
immunoblastic
All organs - all
lymphomas
Hepatocellular
carcinoma
All organs lympho-
immunoblastic
All organs - all
lymphomas
AUC (mg-h/L)
HEDBMDL10(mg/kg-
day)
560.96
555.20
N/A
395.99
397.25
Peak (mg/L)
RED BMDL10
(mg/kg-day)
584.37
578.18
N/A
405.56
407.22
Amount metabolized
(mg/d)a
RED BMDL10
(mg/kg-day)
74.54
71.92
N/A
49.65
51.50b
          "Total metabolized methanol was selected as the preferred dose metric (see discussion in section 5.4.1.3). Before
          applying human PBPK model to obtain these HED BMDL10 estimates, Table E-7 mg/d values were converted to
          human mg/d by multiplying by either: (BWhuman)/4/ (BWrat)y4=(70 kg)y7 (0.33 kg )y<=55.6 for male rats or,
          (BWhuman)yV(BWrat)y4= (70 kg)yV(0.26 kg)y'= 66.5 for female rats.
          bThis value was used in the derivation of the methanol oral cancer slope factor.

          Source: Soffretti et al. (2002a).
                                                       E-15
DRAFT - DO NOT CITE OR QUOTE

-------
    Table E-9. Incidence for neoplasms considered for dose-response modeling
Dose (ppm)
Examined
Adrenal gland
phoechromocytoma
Lung adenoma and
adenocarcinoma
Lung adenoma,
adenocarcinoma,
adenomatosis
Female F344 rats
0
10
100
1,000
50
51
49
51
2
3
2
7








Male F344 rats
0
10
100
1,000
52
50
52
52
Cochran Armitage Trend Test p_values




0.015
1
5
2
T
0.0259
5
6
7
11
0.0415
"Fisher's Exact ^-values < 0.05
Source:  NEDO (1987, 1987/2008b).

    Table E-10. PBPK dose metrics for doses
Dose (ppm)
AUC
(mg-h/L)
Peak (mg/L)
Amount
metabolized (mg)
Female F344 rats
0
10
100
1,000
0.00
3.70
37.53
434.29
0.00
0.19
1.93
22.69
0.00
0.30
2.94
28.96
Male F344 rats
0
10
100
1,000
0
3.70
37.53
433.61
0
0.19
1.93
22.68
0
0.41
4.11
40.19

Source: NEDO (1987, 1987/2008b).
                                         E-16
DRAFT - DO NOT CITE OR QUOTE

-------
               Table E-ll. Benchmark results from multistage quantal dose-response modeling data
               using PBPK dose-metrics
Model
AUC (mg-h/L)
AIC
P-
value
Scaled
residual
at dose
nearest
toBMD
BMD10
BMDL10
Peak (mg/L)
AIC
/)-value
Scaled
residual at
dose
nearest to
BMD
BMD10
BMDL10
Amount metabolized (mg/d)
AIC
/)-value
Scaled
residual
at dose
nearest
to BMD
BMD10
BMDL10
Female F344 rats
Adrenal glands - pheochromocytoma
Multistage (3°)
101.37
0.88
0.000
442.1
217.0
101.37
0.88
0.000
23.1
11.3
101.37
0.88
0.000
29.5
14.6
Male F344 rats
Lung - adenomas and adenocarcinomas
Multistage (3°)
107.99
0.16
0.000
454.9
230.7
107.99
0.16
0.000
23.8
12.1
107.99
0.16
0.001
42.2
21.6
Lung - adenomas, adenocarcinomas and adenomatosis
Multistage (1°)
168.61
0.89
-0.035
378.9
168.8
168.61
0.89
-0.035
19.8
8.83
168.59
0.90
-0.038
34.8
15.6
Source: NEDO (1987, 1987/2008b).
               Table E-12. Application of human PBPK model to derive HECs from BMDLio
               estimates in Table E-ll using multistage quantal modeling

Female F344
rat
Male F344 rat
Adrenal glands -
phoechromocytoma
Lung - adenomas and
adenocarcinomas
Lung - adenomas,
adenocarcinomas and
adenomatosis
AUC (mg-h/L)
HEC BMCL10
(mg/m3)
384.2262
401.96305
316.20336
Peak (mg/L)
HEC BMCL10
(mg/m3)
464.79844
485.15995
386.70859
Amount metabolized (mg/d)a
HEC BMCL10 (mg/m3)
81.85b
101.27
73.12
           a Total metabolized methanol was selected as the preferred dose metric (see discussion in section 5.4.2.3). Before
           applying human PBPK model to obtain these HEC BMCL10 estimates, Table E-7 mg/d values were converted to
           human mg/d by multiplying by either: (BWhuman)/4/ (BWrat)y4=(70 kg)y
-------
Dose (ppm)
Examined
Malignant Lymphoma
Male Swiss Webster mice
Historical untreated controls 3
10
100
1,000
100
25
25
25
8
1
6d
4
aToth et al. (1977); bHinderer et al. (1979); >value = 0.06;
Source: Apaja(1980)
                                                                              < o.05
              Table E-14.  PBPK dose metrics for doses in Apaja (1980)
Daily
Dose
(mg/kg-d)
Weekly
Avg. Dose
(mg/kg-d)
Body
Weight
(kg)
AUC
(mg-h/L)
Peak (mg/L)
Amount
metabolized (mg)
Female Swiss Webster mice
0
560.01
1000.00
2099.98
0
479.99
857.13
1800.00
0.040
0.040
0.040
0.040
0
484.86
3466.74
19503.87
0
88.40
383.10
1462.07
0
18.38
28.45
39.18
Male Swiss Webster mice
0
550.00
970.00
1800.00
0
471.42
831.39
1542.86
0.045
0.045
0.045
0.045
0
500.63
3405.26
14998.60
0
89.77
373.72
1162.69
0
20.29
31.15
41.86
                   Source: Apaja (1980).

              Table E-15.  Benchmark results from Multistage-cancer dose-response modeling data
              for malignant lymphoma in Swiss Webster mice (Apaja, 1980) using PBPK dose-
              metrics
Gender
Female mice3
Male mice3
AUC (mg-h/L)
AIC
288.91
122.09
P-
value
0.32
0.083
Scaled
residual
at dose
nearest
toBMD
1.354
-0.883
BMD10
5808.3
11164.2
BMDL10
2957.6
4343.04
Peak (mg/L)
AIC
288.42
121.58
/)-value
0.43
0.12
Scaled
residual at
dose
nearest to
BMD
1.090
-0.641
BMD10
428.5
798.3
BMDL10
225.1
338.9
Amount metabolized (mg/d)
AIC
288.07
120.97
7>-value
0.55
0.21
Scaled
residual
at dose
nearest
to BMD
-0.937
1.309
BMD10
22.7
35.8
BMDL10
10.1
18.3
3Multistage-cancer (1°) used for AUC and Peak metrics; Multistage-cancer (2°) used for Amount metabolized metric
Source: Apaja (1980).

      1
                                                    E-18
                                           DRAFT - DO NOT CITE OR QUOTE

-------
                  Table E-16.  Application of human PBPK model to derive HEDs from BMDLio
                  estimates of Table E-15,  Multistage (1°) modeling of malignant lymphoma in Swiss
                  mice (Apaja, 1980)  using PBPK dose metrics

Female mice
Male mice
AUC (mg-h/L)
RED BMDL10 (mg/kg-day)
248
269
Peak (mg/L)
RED BMDL10 (mg/kg-
day)
281
305
Amount metabolized (mg/d)a
RED BMDL10
(mg/kg-day)
9.7
14.7
             "Before applying human PBPK model Table 11 mg/day values were converted to human mg/day by multiplying by
             either: (BWhuman)y7 (BWratf=(70 kgf/ (0.33 kg f=55.6 for male rats or, (BWhuman)yV(BWratf = (70 kgf/(0.26 kg f
             = 66.5 for female rats.
                  Table E-17.  Benchmark results for all tumor types using BMDS 2.1 multistage
                  "background dose" models and PBPK dose-metrics
     Model
                           AUC (mg-h/L)
                 AIC
                       P-
                      value
Background
   Dose
                                      BMD10
BMDL10
                                                                 Peak (mg/L)
                                                     AIC
 p-
value
Background
   Dose
BMD10
BMDL10
                                                                    Amount metabolized (mg/d)
                                                                                           AIC
 p-
value
Background
   Dose
BMD10  BMDL10
                                        Female Sprague-Dawley rats (Soffritti et al., 2002a)
All organs lympho-immunoblastic
Multistage (1°)   361.99 0.045    168476   112790   49771   359.93  0.14    7480.0     5034.3   2240.3   357.87  0.34
                                                                                                         169.0
                                                                                                                 139.8    77.2
All organs - all lymphomas
Multistage (1°)   372.78 0.075    179247   111621   48790   372.72  0.077   7961.1     4982.2   2196.3   370.57  0.18
                                                                                                         175.1
                                                                                                                 134.4    74.1
                                         Male Sprague-Dawley rats (Soffritti et al., 2002a)
Hepatocellular carcinoma
Multistage (1°)    97.52   o.oo
                                     Failed8
                                                     73.19  0.37
                                                                          Failed"
                                                                                           72.55  0.36
                                                                                                                Failed8
All organs lympho-immunoblastic
Multistage (1°)   455.i1  0.18   85497.6    36501   21901   454.97  0.20    3781.3     1625.7    979.1    453.98  0.30
                                                                                                         190.2
                                                                                                                 97.8     63.2
All organs - all lymphomas
Multistage (1°)   468.09  0.15   96684.3    37822   22195   467.98  0.15    4285.6     1686.5    993.1    467.35  0.20
                                                                                                         221.2
                                                                                                                 101.7    63.9
                                              Female F344 rats (NEDO, 1985/2008b)
Adrenal glands - pheochromocytoma
Multistage (2°)   ioi.46  0.84    553.1     450.9   214.8   103.45  0.56
                                                                   28.9
                                                                            23.6
                                                                                    11.2    101.47  0.83
                                                                                                         36.7
                                                                                                                 30.1
                                                                                                                         14.4
                                               Male F344 rats (NEDO, 1985/2008b)
Lung - adenomas and adenocarcinomas
Multistage (1°)   108.15 | 0.14 |   248.16    507.9 I  226.0  | 108.15  0.14     13.0    |  26.6   |  11.8  |  108.19  0.14
                                                                                                         23.0
                                                                                                                 47.3     21.0
Lung - adenomas, adenocarcinomas amd adenomatosis
Multistage (1°)   168.61  0.89    430.38    378.9   168.8   168.61  0.89     22.5
                                                                             19.8
                                                                                     8.8    168.59  0.90
                                                                                                         39.3
                                                                                                                 34.8
                                                                                                                         15.6
                                            Female Swiss Webster mice (Apaja, 1989)
Malignant lymphomas
Multistage (1°)   288.91 | 0.32 |   11933.3    5808  I  2958  | 288.42  0.43     863.1   |  428.5  |  225.1  |  289.00  0.37
                                                                                                                  16.8
                                                                                                                          9.1
                                             Male Swiss Webster mice (Apaja, 1989)
Multistage (1°)   122.09 0.083    9573.9    11164   4343.0   121.58  0.12     646.9     798.3    338.9   121.30  0.20
                                                                                                         28.6
                                                                                                                 38.1     17.6
aBMD computation failed. BMD is larger than three times maximum input doses.
Source:  NEDO (1987, 1987/2008b).
                                                             E-19
                                                                               DRAFT - DO NOT CITE OR QUOTE

-------
                                       Female Rats Survival
                                         60     80     100
                                         Time on Study (weeks)
                                                            120
                                                                   140
                                                                         160
                              -Control •
                                       -500 ppm •
-5000 ppm
20000 ppm
Figure E-l.  Female rat survival.
   Source:  Soffretti et al. (2002a).
                                      Male Rats Survival
                               30    45    60    75    90    105   120   135   150
                                        Time on Study (weeks)
                               -Control	500 ppm	5000 ppm	20000 ppm
Figure E-2.  Male rats survival.
   Source:  Soffretti et al. (2002a).
                                           E-20
            DRAFT - DO NOT CITE OR QUOTE

-------
                  14:2701/11/2007
                                  Female Rats - Multistage Weibull Model (t= 105 weeks)
                                                         Time-to-Tumor Weibull
                                                   Observed (90%% Binomial limits)
                                            1000          1500

                                          Animal Dose (mg/kg-day)
Figure E-3. Female - Lymphomas lympho-immunoblastic - Multistage Weibull Model
-Approach 1.
   Source:  Soffretti et al. (2002a).
                  14:0201/11/2007
                                  Female Rats - Multistage Weibull Model (1= 105 weeks)
                                                         Time-to-Tumor Weibull
                                                   Observed (90%% Binomial limits)
                                            1000          1500

                                          Animal Dose (mg/kg-day)
Figure E-4. Female - All lymphomas - Multistage Weibull Model - Approach 1.
   Source:  Soffretti et al. (2002a).
                                            E-21
DRAFT - DO NOT CITE OR QUOTE

-------
                  15:1905/21/2008
                                   Risk Assessment: 2 ; Chemical: Methanol; Sex: Male
                                                          Time-to-TumorWeibull
                                                    Observed (90%% Binomial limits)
                      0     200    400   600    800    1000    1200   1400   1600   1800

                                            Dose (mg/kg/day)
Figure E-5. Male - Hepatocellular carcinoma - Multistage Weibull Model - Approach
1.
   Source:  Soffretti et al. (2002a).
                      16:0301/11/2007
                                       Male Rats- Multistage Weibull Model (t= 105 weeks)
                                                              Time-to-Tumor Weibull
                                                        Observed (90%% Binomial limits)
                               200    400   600    800    1000    1200   1400   1600   1800

                                              Animal Dose (mg/kg-day)
Figure E-6. Male - Lymphomas lympho-immunoblastic - Multistage Weibull Model
Approach 1.

   Source: Soffretti et al. (2002a).
                                            E-22
DRAFT - DO NOT CITE OR QUOTE

-------
                 16:0401/11/2007
                                  Male Rats- Multistage Weibull Model (t= 105 weeks)
                                                        Time-to-Tumor Weibull
                                                  Observed (90%% Binomial limits)
                          200    400    600    800   1000   1200    1400   1600   1800
                                         Animal Dose (mg/kg-day)
Figure E-7.  Male - All lymphomas - Multistage Weibull Model - Approach 1.
   Source:  Soffretti et al. (2002a).
                                     Female Rats All Lymphomas Combined
                                   Kaplan Meier Plots and Multistage Weibull Model
Figure E-8.  Female rats -All lymphomas time-to-tumor model fit and Kaplan Meier
curves.
   Source:  Soffretti et al. (2002a).
                                           E-23
DRAFT - DO NOT CITE OR QUOTE

-------
                                   Male Rats All Lymphomas Combined

                                 Kaplan Meier Plots and Multistage Weibull Model
Figure E-9.  Male rats -All lymphomas time-to-tumor model fit and Kaplan Meier

curves.

   Source:  Soffretti et al. (2002a).
                             Multistage Cancer Model with 0.95 Confidence Level
       .3
       V
       £
               0.1
         08:55 09/03 2009
                                      Multistage Cancer
                                     Linear extrapolation
                                                                      200
Figure E-10. Male rats-All lymphomas; dose = amount metabolized (mg/day); 1°
multistage model.

   Source:  Soffretti et al. (2002a).
                                        E-24
DRAFT - DO NOT CITE OR QUOTE

-------
                •S 0.7
                  0.5
                I.
                  0.2
                    50
                                     Female Rats Survival
                             60
                                      70
                                       Time on Study (weeks)
                              -Control •
                                      - lOppm -
-lOOppm
1000 ppm
Figure E-ll.  Female rat survival.
   Source: NEDO (1987, 1987/2008b).
                   i
                  0.9
                  0.8 -
                  0.7 -
                  0.6 -
                  0.5 -
                  0.4 -
                  0.3 -
                  0.2 -
                  0.1 -
                    50
                                    Male Rats Survival
                                  65
                                                             95
                                       Time on Study (weeks)
                               -Control •
                                        - lOppm -
 -100 ppm   1000 ppm
Figure E-12. Male rat survival.
   Source: NEDO (1987, 1987/2008b).
                                         E-25
           DRAFT - DO NOT CITE OR QUOTE

-------
                    08:58 09/03 2009
                                  Multistage Cancer Model with 0.95 Confidence Level
                                        Multistage Cancer
                                        Linear extrapolation
                                              15

                                             dose
Figure E-13. Female rats- pheochromocytomas; dose = amount metabolized (mg/d); 3°
multistage model.

   Source: NEDO (1987, 1987/2008b)
                                  Multistage Cancer Model with 0.95 Confidence Level
                                       Multistage Cancer
                                       Linear extrapolation
                     09:58 09/03 2009
                                         15    20   25   30    35   40

                                             dose
                                  Multistage Cancer Model with 0.95 Confidence Level
                                       Multistage Cancer
                                       Linear extrapolation
                                    10   15   20    25   30   35   40
                    09:56 09/03 2009
Figure E-14. Plots for female (top; p= 0.55) and male (bottom; p= 0.21) mice -
malignant lymphoma; dose=amount metabolized (mg/d); 2° multistage model.

   Source: Apaja(1980)
                                         E-26
DRAFT - DO NOT CITE OR QUOTE

-------