Skip to main content

Abstract

The purpose of this chapter is to review experimental studies, especially those involving combustion chemistry and toxicity test methods, in order to establish the basis and validation for material toxicity and toxic hazard calculations and to obtain yield data for input to fire dynamics and toxicity calculations. The review covers several major topics, including the following:

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 869.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 1,099.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. D.T. Gottuk and B.Y. Lattimer, “Effect of Combustion Conditions on Species Production.” in SFPE Handbook of Fire Protection Engineering, 5th ed. (M. J. Hurley et al., eds.), Springer, (2015).

    Google Scholar 

  2. Khan, M. et al., “Combustion Characteristics of Materials and Generation of Fire Products,” SFPE Handbook of Fire Protection Engineering, 5th ed. (M. J. Hurley et al., eds.), Springer, (2015).

    Google Scholar 

  3. D.A.Purser and J.L.McAllister. “Assessment of Hazards to Occupants from Smoke, Toxic Gases and Heat” SFPE Handbook of Fire Protection Engineering, 5th ed. (M. J. Hurley et al., eds.), Springer, (2015).

    Google Scholar 

  4. ISO 13571 Life-threatening components of fire –Guidelines for the estimation of time to compromised tenability in fires. Second edition International Organization for Standardization, Geneva, Switzerland. 2012.

    Google Scholar 

  5. J.H. Petajan, K.L. Voorhees, S.C. Packham, R.C. Baldwin, I.N. Einhorn, M.L. Grunnet, B.G. Dinger, and M.M. Birky, “Extreme Toxicity from Combustion Products of a Fire-Retarded Polyurethane Foam,” Science, 187, pp. 742–744 (1975).

    Article  Google Scholar 

  6. B.C. Levin, A.J. Fowell, M.M. Birky, M. Paabo, S. Stolte, and D. Malek, “Further Development of a Test Method for the Acute Inhalation Toxicity of Combustion Products,” NBSIR 82-2532, National Bureau of Standards, Washington, DC (1982).

    Google Scholar 

  7. R.C. Anderson, P.A. Croce, F.G. Feeley, and J.D. Sakura, “Study to Assess the Feasibility of Incorporating Combustion Toxicity Requirements into Building Materials and Furnishing Codes of New York State,” Reference 88712, Arthur D. Little, Cambridge, MA (1983).

    Google Scholar 

  8. C.J. Hilado, “Screening Materials for Relative Toxicity in Fire Situations,” Modern Plastics, pp. 64–68 (July, 1977).

    Google Scholar 

  9. D.A. Purser,. “Application of human and animal exposure studies to human fire safety”. in Fire Toxicity, A.A. Stec and T.R. Hull (eds) Woodhead, Cambridge, 2010. Chapter 8 pp 283–345.

    Google Scholar 

  10. D.A. Purser, Validation of additive models for lethal toxicity of fire effluent mixtures. Polymer Degradation and Stability 97 2552–2561 (2012)

    Article  Google Scholar 

  11. Controlled Equivalence Ratio Method for the Determination of Hazardous Components of Fire Effluents,” ISO/TS 19700, International Organization for Standardization, Geneva (2007).

    Google Scholar 

  12. D.A. Purser, P.J. Fardell, J. Rowley, S. Vollam, and B. Bridgeman “An Improved Tube Furnace Method for the Generation and Measurement of Toxic Combustion Products Under a Wide Range of Fire Conditions,” in Proceedings of Flame Retardants 1994 Conference, Interscience Communications, London, pp. 263–274 (1994).

    Google Scholar 

  13. J.A. Purser, D.A. Purser, A.A. Stec, C. Moffatt, T.R. Hull, J.Z. Su, M. Bijloos, P. Blomqvist. Repeatability and reproducibility of the ISO/TS 19700 steady state tube furnace. Fire Safety Journal 55, 22–34 (2013)

    Google Scholar 

  14. “Guidelines for Methodology for Assessing the Fire Threat to People,” ISO 19706, ISO, Geneva, Switzerland (2007).

    Google Scholar 

  15. R.A. Anderson, A.A. Watson, and W.A. Harland, “Fire Deaths in the Glasgow Area: 1. General Conclusions and Pathology,” Medicine, Science and the Law, 21, pp. 175–183 (1981).

    Article  Google Scholar 

  16. H.L. Kaplan, A.F. Grand, and G.E. Hartzell, Combustion Toxicology: Principles and Test Methods, Technomic, Lancaster, PA (1983).

    Google Scholar 

  17. D.A. Purser and W.D. Woolley, “Biological Studies of Combustion Atmospheres,” Journal of Fire Sciences, 1, pp. 118–145 (1983).

    Article  Google Scholar 

  18. W.D. Woolley and P.J. Fardell, “Basic Aspects of Combustion Toxicology,” Fire Safety Journal, 5, p. 29 (1982).

    Article  Google Scholar 

  19. F. Haber, Funf Vortrange aus den jaren 1920–1923, Verlag von Julius Springer, Berlin (1924).

    Google Scholar 

  20. D.A. Purser, Chapter 4: Asphyxiant components of fire effluents. In “Fire toxicity” Eds A. Stec and R. Hull. Woodhead, Cambridge UK, 2010 pp 118–198

    Google Scholar 

  21. D.A. Purser and M. Kuipers, M. (2004) Interactions between buildings, fire and occupant behaviour using a relational database created from incident investigations and interviews. 3rd International Symposium on Human Behaviour in Fire. Europa Hotel, Belfast, 1–3rd September 2004. Proceedings pp. 443–456 Interscience Communications, London UK.

    Google Scholar 

  22. G.L. Nelson, “Carbon Monoxide and Fire Toxicity: A Review and Analysis of Recent Work, “Fire Technology, 34, pp. 38–58 (1998).

    Article  Google Scholar 

  23. R.A. Anderson, I. Thompson, and W.A. Harland, “The Importance of Cyanide and Organic Nitriles in Fire Fatalities,” Fire and Materials, 3, pp. 91–99 (1979).

    Article  Google Scholar 

  24. J.L. McAllister, R.J. Roby, B. Levine, and D. Purser, Stability of cyanide in cadavers and in postmortem stored tissue specimens: A review. J. Analytical Toxicology. 32, 1–9(2008)

    Google Scholar 

  25. D.A. Purser, P. Grimshaw, and K.R. Berrill, “Intoxication by Cyanide in Fires: A Study in Monkeys Using Polyacrylonitrile,” Archives of Environmental Health, 39, pp. 394–400 (1984).

    Article  Google Scholar 

  26. D.A. Purser, “Determination of Blood Cyanide and Its Role in Producing Incapacitation in Fire Victims,” Royal Society of Chemistry Meeting, Huntingdon (1984).

    Google Scholar 

  27. D.A. Purser, “Toxic Product Yield and Hazard Assessment for Fully Enclosed Design Fires Involving Fire Retarded Materials,” Polymer International, 47, pp. 1232–1255 (2000).

    Article  Google Scholar 

  28. D.A. Purser, J.A. Rowley, P.J. Fardell, and M. Bensilum, “Fully Enclosed Design Fires for Hazard Assessment in Relation to Yields of Carbon Monoxide and Hydrogen Cyanide,” Interflam’99. Eighth International Fire Science and Engineering Conference, Edinburgh, Proceedings pp. 1163–1169, Interscience Communications, London (June–July 1999).

    Google Scholar 

  29. F.J. Baud, P. Barriot, V. Toffis, et al., “Elevated Blood Cyanide Concentrations in Victims of Smoke Inhalation,” New England Journal of Medicine, 325, pp. 1761–1766 (1991).

    Article  Google Scholar 

  30. King, D.F. (1988) Aircraft Accident Report 8/88, UK Department if Transport. Air Accidents Investigation Branch. HMSO. London, UK.

    Google Scholar 

  31. D.A. Purser and J.A. Purser, HCN yields and fate of fuel nitrogen for materials under different combustion conditions in the ISO 19700 tube furnace. Fire Safety Science – Proceedings of the ninth international symposium. 2008. pp 1117–11128. International Association for Fire Safety Science.

    Google Scholar 

  32. D.A. Purser and P. Buckley, “Lung Irritance and Inflammation During and After Exposure to Thermal Decomposition Products from Polymeric Materials,” Medicine Science and the Law, 23, pp. 142–150 (1983).

    Article  Google Scholar 

  33. D.A. Purser and K.R. Berrill, “Effects of Carbon Monoxide on Behaviour in Monkeys in Relation to Human Fire Hazard,” Archives of Environmental Health, 38, pp. 308–315 (1983).

    Article  Google Scholar 

  34. D.A. Purser, “A Bioassay Model for Testing the Incapacitating Effects of Exposure to Combustion Product Atmospheres Using Cynomolgus Monkeys,” Journal of Fire Sciences, 2, pp. 20–26 (1984).

    Article  Google Scholar 

  35. D.A. Purser and P. Grimshaw. (1984) The incapacitative effects of exposure to the thermal decomposition products of polyurethane foams. Fire and Materials. 8, 10–16.

    Article  Google Scholar 

  36. Kaplan H. L., Grand, A.F., Switzer, W.G., Mitchell, D.S., Rogers, W.R. and Hartzell, G.E. Effects of Combustion Gases On Escape Performance of the Baboon and the Rat Journal of Fire Sciences, Vol. 3, No. 4, 228–244 (1985) DOI: 10.1177/073490418500300401

    Google Scholar 

  37. W.D. Woolley, S.A. Ames, and P.J. Fardell, “Chemical Aspects of Combustion Toxicology of Fires,” Fire Materials, 3, pp. 110–120 (1979).

    Google Scholar 

  38. D.A. Purser, A.A Stec and T.R. Hull, Chapter 2: Fire scenarios and combustion conditions. In “Fire toxicity” Eds A. Stec and R. Hull. Woodhead, Cambridge UK, 2010 pp 26–50

    Google Scholar 

  39. D.A. Purser, A.A. Stec and T.R. Hull, Chapter 14: Effects of material and fire conditions on toxic product yields. In “Fire toxicity” Eds A. Stec and R. Hull. Woodhead, Cambridge UK, 2010 pp 515–540.

    Google Scholar 

  40. V. Babrauskas, B.C. Levin, R.G. Gann, et al., Toxic Potency Measurement for Fire Hazard Analysis, NIST Special Publication 827 (1991).

    Google Scholar 

  41. DIN, “Producing Thermal Decomposition Products from Materials in an Air Stream and Their Toxicological Testing,” DIN 53 436, Deutsches Institut für Normung, Berlin, Germany.

    Google Scholar 

  42. H. Klimisch, H.W. Hollander, and J. Thyssen, “Comparative Measurements of the Toxicity to Laboratory Animals of Products of Thermal Decomposition Generated by the Method of DIN 53 436,” Journal of Combustion Toxicology, 7, pp. 209–230 (1980).

    Google Scholar 

  43. Jounay, J.M., Boudene, C. and Truhout, R. The physiogram as a method for the evaluation of combustion products in controlled ventilation experiments. Polymer Conference Salt Lake City, 1976.

    Google Scholar 

  44. Purser, D.A. (1992) The evolution of toxic effluents in fires and the assessment of toxic hazard. Toxicology Letters, 64/65, 247–255. doi:10.1016/0378-4274(92)90196-Q

    Article  Google Scholar 

  45. Determination of the Lethal Toxic Potency of Fire Effluents,” ISO 13344 (1996).

    Google Scholar 

  46. Levin, B.C.(1996) New research avenues in toxicology: 7-gas N-gas model, toxicant suppressants and genetic toxicology. Toxicology, 115: 89–106. doi:10.1016/S0300-483X(96)03497-X

    Article  Google Scholar 

  47. Hartzell, G.E., Grand A.F and Switzer, W.G (1987) Modeling of Toxicological Effects of Fire Gases: VI. Further Studies on the Toxicity of Smoke Containing Hydrogen Chloride Journal of Fire Sciences, 5:368–391 DOI: 10.1177/073490418700500602

    Google Scholar 

  48. B. Levin, J. Gurman, M. Paabo, L. Baier and T. Holt. Toxicological effects of different time exposures to fire gases: carbon monoxide or hydrogen cyanide or to carbon monoxide combined with hydrogen cyanide or carbon dioxide. Proceedings of the 9th Joint Panel Meeting of the UNJR Panel on Fire Research and Safety, US National Bureau of Standards, Gaithersburg. MD. Report NBSIR 88–3753, p 368, 1988.

    Google Scholar 

  49. B.C. Levin, M. Paabo, J.L. Gurman, and S.C. Harris, “Effects of Exposure to Single or Multiple Combinations of the Predominant Toxic Gases and Low-Oxygen Atmospheres Produced in Fires,” Fundamental and Applied Toxicology, 9, pp. 236–250 (1987). doi:10.1016/0272-0590(87)90046-7

    Article  Google Scholar 

  50. J.A. Pauluhn Retrospective Analysis of Predicted and Observed Smoke Lethal Toxic Potency Values. J. Fire Science 11 109–130 (1993).

    Article  Google Scholar 

  51. “Documentation of the Threshold Limit Values for Substances in Workroom Air,” American Conference of Governmental Industrial Hygienists, Cincinnati (1980).

    Google Scholar 

  52. F.W. Beswick, P. Holland, and K.H. Kemp, “Acute Effects of Exposure to Orthochlorobenzylidene Malonitrile (CS) and the Development of Tolerance,” British Journal of Industrial Medicine, 29, pp. 298–306 (1972).

    Google Scholar 

  53. C.L. Punte, J.T. Weimer, T.A. Ballard, and J.L. Wilding, “Toxicologic studies on o-Chlorobenzylidene Malononitrile,” Toxicology and Applied Pharmacology, 4, pp. 656–662 (1962).

    Article  Google Scholar 

  54. B. Ballantyne and S. Calloway, “Inhalation Toxicology and Pathology of Animals Exposed to o-Chlorobenzylidene Malonitrile,” Medicine, Science and the Law, 12, pp. 43–65 (1972).

    Google Scholar 

  55. Registry of Toxic Effects of Chemical Substances, National Institute for Occupational Safety and Health, Washington, DC (1982).

    Google Scholar 

  56. L. Kane, C.S. Barrow, and Y. Alarie, “A Short-Term Test to Predict Acceptable Levels of Exposure to Airborne Sensory Irritants,” American Industrial Hygiene Association Journal, 40, pp. 207–209 (1979).

    Article  Google Scholar 

  57. H.L. Kaplan, A.F. Grand, W.R. Rogers, W.G. Switzer, and G.C. Hartzell, “Combustion Gases in Postcrash Aircraft Fires,” DOT/FAA/CT-84/16, Federal Aviation Administration, Washington, DC (1984

    Google Scholar 

  58. Y. Alarie, Proceedings of the Inhalation Toxicology Symposium, Upjohn Company, Ann Arbor Science (The Butterworth Group), Ann Arbor, MI (1980).

    Google Scholar 

  59. H. Salem and H. Cullumbine, “Inhalation Toxicities of Some Aldehydes,” Toxicology and Applied Pharmacology, 2, pp. 183–187 (1960).

    Article  Google Scholar 

  60. T.J. Cole, J.E. Cotes, G.R. Johnson, H. deV. Martin, J.W. Reed, and M.J. Saunders, “Ventilation, Cardiac Frequency and Pattern of Breathing During Exercise in Men Exposed to o-Chlorobenzylidene Malonitrile (CS) and Ammonia Gas in Low Concentrations,” Quarterly Journal of Experimental Physiology, 62, pp. 341–351 (1977).

    Google Scholar 

  61. Y. Alarie, “Bioassay for Evaluating the Potency of Airborne Sensory Irritants and Predicting Acceptable Levels of Exposure in Man,” Food and Cosmetics Toxicology, 19, pp. 623–626 (1981).

    Article  Google Scholar 

  62. D.A. Purser, “Behavioural Impairment in Smoke Environments,” Toxicology, 115, pp. 25–40 (1996).

    Article  Google Scholar 

  63. “British Standard Code of Practice for Assessment of Hazard to Life and Health from Fire. Part 2: Guidance on Methods for the Quantification of Hazards to Life and Health and Estimation of Time to Incapacitation and Death in Fires,” BS 7899-2, British Standards, London (1999).

    Google Scholar 

  64. D.A. Purser, “The Harmonization of Toxic Potency Data for Materials Obtained from Small and Large Scale Tests and Their Use in Calculations for the Prediction of Toxic Hazard in Fire,” Proceedings of First International Fire and Materials Conference, Interscience Communications, London, pp. 179–200 (1992).

    Google Scholar 

  65. D.A. Purser, “Recent Developments in Understanding the Toxicity of PTFE Thermal Decomposition Products,” Fire and Materials, 16, pp. 67–75 (1992).

    Article  Google Scholar 

  66. Purser, D.A and Purser J.A. The potential for including fire chemistry and toxicity in fire safety engineering. BRE Project Report No. 202804. 26th March 2004. Building Research Establishment. Garston Watford UK.

    Google Scholar 

  67. D.A. Purser, Toxicity of fire retardants in relation to life safety and environmental hazards. In: Fire Retardant Materials. Ed. A.R. Horrocks and D. Price. Chapter 3 pp. 69–127.(2001) Woodhead Publishing Ltd, Cambridge UK.

    Chapter  Google Scholar 

  68. D.A. Purser, “Recent Developments in the Use of a Tube Furnace Method for Evaluating Toxic Product Yields under a Range of Fire Conditions,” Fire Protection Research Association 2000 Fire Risk and Hazard Research Application Symposium, Atlantic City (June 28–30, 2000).

    Google Scholar 

  69. T.R. Hull, J.M. Carmen, and D.A. Purser, “Prediction of CO2/CO Ratios of Underventilated Polymer Fires,” Polymer International, 49, pp. 1259–1265 (2000).

    Article  Google Scholar 

  70. W.M. Pitts, “The Global Equivalence Ratio Concept and the Formation Mechanisms of Carbon Monoxide in Enclosure Fires,” Progress in Energy and Combustion Science, 21, pp. 197–237 (1995).

    Article  Google Scholar 

  71. D.A. Purser, The application of exposure concentration and dose to evaluation of the effects of irritants as components of fire hazard. Interflam 2007. 3–5th September 2007 Royal Holloway College, Egham, UK. Proceeding pp. 1033–1041. Interscience Communications, Greenwich, UK.

    Google Scholar 

  72. Purser D.A, Chapter 3: Hazards from smoke and irritants. In “Fire toxicity” Eds A. Stec and R. Hull. Woodhead, Cambridge UK, 2010 pp 51–117.

    Google Scholar 

  73. D. Campbell, Respiratory Tract Trauma in Burned Patients, Colloquium, Borehamwood, UK (1985).

    Google Scholar 

  74. C.L. Punte, E.J. Owens, and P.J. Gutentag, “Exposures to ortho-Chlorobenzylidene Malonitrile,” Archives of Environmental Health, 6, pp. 366–374 (1963).

    Article  Google Scholar 

  75. T. Jin, “Studies of Emotional Instability in Smoke from Fires,” Journal of Fire and Flammability, 12, pp. 130–142 (1981).

    Google Scholar 

  76. D.J. Rasbash, Fire International, 5, 40, p. 30 (1975).

    Google Scholar 

  77. Y. Alarie, “Sensory Irritation by Airborne Chemicals,” CRC Critical Reviews in Toxicology, 2, p. 299 (1973).

    Article  Google Scholar 

  78. D. Canter, Studies of Human Behavior in Fire: Empirical Results and Their Implications for Education and Design, University of Surrey, UK (1983).

    Google Scholar 

  79. R. Melzack and P.D. Wall, “Pain Mechanisms: A New Theory,” Science, 150, pp. 971–979 (1965).

    Article  Google Scholar 

  80. D.G. Clark, S. Buch, J.E. Doe, H. Frith, and D.H. Pullinger, “Bronchopulmonary Function: Report of the Main Working Party,” Pharmacology and Therapeutics, 5, pp. 149–179 (1979).

    Article  Google Scholar 

  81. C.S. Barrow, H. Lucia, M.F. Stock, and M.F. and Y. Alarie, “Development of Methodologies to Assess the Relative Hazards from Thermal Decomposition Products of Polymeric Materials,” American Industrial Hygiene Association Journal, 40, pp. 408–423 (1979).

    Google Scholar 

  82. Y. Alarie and R.C. Anderson, “Toxicologic and Acute Lethal Hazard Evaluation of Thermal Decomposition Products of Synthetic and Natural Polymers,” Toxicology and Applied Pharmacology, 51, pp. 341–362 (1979).

    Article  Google Scholar 

  83. W.J. Potts and T.S. Lederer, “A Method for Comparative Testing of Smoke Toxicity,” Journal of Combustion Toxicology, 4, pp. 114–162 (1977).

    Google Scholar 

  84. T. Jin, “Visibility through Fire Smoke—Part 5, Allowable Smoke Density for Escape from Fire.” Report No. 42, Fire Research Institute of Japan, p. 12 (1976).

    Google Scholar 

  85. D.A. Purser, “Interactions Among Carbon Monoxide, Hydrogen Cyanide, Low Oxygen Hypoxia, Carbon Dioxide and Inhaled Irritant Gases,” in Carbon Monoxide Toxicity Chapter 7 pp. 157–192. D.G. Penney ed, CRC Press Boca Raton (2000)

    Google Scholar 

  86. M.M. Hirschler and D.A. Purser, “Irritancy of the Smoke (Non-Flaming Mode) from Materials Used for Coating Wire and Cable Products, Both in the Presence and Absence of Halogens in their Chemical Composition,” Fire and Materials, 17, pp. 7–20 (1993).

    Article  Google Scholar 

  87. D.A. Purser, “Modeling Time to Incapacitation and Death from Toxic and Physical Hazards in Aircraft Fires,” in Conference Proceedings No. 467, Aircraft Fire Safety, NATO-AGARD, Sintra, Portugal, pp. 41-1–41-13 (May 22–26, 1989).

    Google Scholar 

  88. Emergency Response Planning Guidelines, American Industrial Hygiene Association, Akron, OH (1989).

    Google Scholar 

  89. NAC/AEGL, “Acute Exposure Guidelines for Selected Airborne Chemicals,” National Advisory Committee for Acute Exposure Guideline Levels for Hazardous Substances, Subcommittee on Acute Exposure Guideline Levels, Committee on Toxicology, Board on Environmental Studies and Toxicology, National Research Council of the National Academies, National Academies Press, Washington, DC (2004). Note: Reviews of new substances are being added on a continuing basis.

    Google Scholar 

  90. D.A. Purser, “ASET and RSET: Addressing Some Issues in Relation to Occupant Behaviour and Tenability,” in Proceedings of the 7th International Symposium on Fire Safety Science, International Association for Fire Safety Science, Boston, MA, pp. 91–102 (2003).

    Google Scholar 

  91. T.R. Hull, D. Price, J.M. Carman, and D. Purser, “Studies of Carbon/Oxygen Chemistry Under Different Fire Conditions,” in Proceedings of the Interflam’99 Conference, Vol. 2, pp. 189–199, Interscience Communications Ltd., London, UK (1999).

    Google Scholar 

  92. M. Carman, D.A. Purser, T.R. Hull, D. Price, and G.J. Milnes, “Experimental Parameters Affecting the Performance of the Purser Furnace—A Laboratory Scale Experiment for a Range of Controlled Real Fire Conditions,” Fire Retardant Polymers, 7th European Conference, in Polymer International, 49, pp. 1256–1258 (2000).

    Google Scholar 

  93. P. Blomqvist, “A Small-Scale Controlled Equivalence Ratio Tube Furnace Method—Experience of the Method and the Link to Large-Scale Fires,” in Proceedings of the Interflam 2007 Conference, London, UK, pp. 391–402 (2007).

    Google Scholar 

  94. “Tube Furnace Method for the Determination of Toxic Product Yields in Fire Effluents,” BS7900 British Standards Institution, London, UK (2003).

    Google Scholar 

  95. NFX 70-100-1, “Methods for Analysing Gases Stemming from Thermal Degradation,” AFNOR, Paris, France (2001).

    Google Scholar 

  96. NFPA 287, Standard Test Methods for Measurement of Flammability of Materials in Cleanrooms Using a Fire Propagation Apparatus (FPA), National Fire Protection Association International, Quincy, MA (2007).

    Google Scholar 

  97. “Reaction-to-Fire Tests—Heat Release, Smoke Production, and Mass Loss Rate—Part 1: Heat Release (cone calorimeter method),” ISO 5660-1, ISO, Geneva, Switzerland (2002).

    Google Scholar 

  98. ASTM E662, Standard Test Method for Specific Optical Density of Smoke Generated by Solid Materials, ASTM International, West Conshohocken, PA.

    Google Scholar 

  99. IMO, FTP Code: International Code for Application of Fire Test Procedures, Resolution MSC 61(67), International Maritime Organization, London, UK (1998).

    Google Scholar 

  100. “Code of Practice for Fire Precautions in the Design and Construction of Passenger Carrying Trains,” BS 68553, British Standards Institution, London, UK (1999).

    Google Scholar 

  101. “Toxicity Testing of Fire Effluents—Part 5. Prediction of Toxic Effects of Fire Effluents,” ISO/IEC TR 9122-5 (1993).

    Google Scholar 

  102. “Code of Practice for Assessment of Hazard to Life and Health from Fire—Part 2. Guidance on Methods for the Quantification of Hazards to Life and Health and Estimation of Time to Incapacitation and Death in Fires, B57899-2 (1999).

    Google Scholar 

  103. G. Kimmerle and F.C. Prager, “The Relative Toxicity of Pyrolysis Products. Part II. Polyisocyanate-Based Foam Materials,” Journal of Combustion Toxicology, 7, pp. 54–69 (1980).

    Google Scholar 

  104. V. Babrauskas, “Development of the Cone Calorimeter—A Bench Scale Heat Release Rate Apparatus Based on Oxygen Consumption,” NBSIR 82-2611, National Bureau of Standards, Washington, DC (1982).

    Google Scholar 

  105. “Toxicity Testing of Fire Effluents—Part 4, The Fire Model.” ISO/IEC TR 9122-4, International Organization for Standardization, Geneva, Switzerland (1993).

    Google Scholar 

  106. G.E. Hartzell, W.G. Switzer, and D.N. Priest, “Modeling of Toxicological Effects of Fire Gases: V Mathematical Modelling of Intoxication of Rats by Combined Carbon Monoxide and Hydrogen Cyanide Atmospheres,” Journal of Fire Sciences, 3, pp. 330–342 (1985).

    Article  Google Scholar 

  107. G.E. Hartzell, S.C. Packham, A.F. Grand, and W.G. Switzer, “Modeling of Toxicological Effects of Fire Gases: III. Quantification of Post-Exposure Lethality of Rats from Exposure to HCI Atmospheres,” Journal of Fire Sciences, 3, pp. 196–207 (1985).

    Google Scholar 

  108. W.A. Burgess, R.D. Trietman, and A. Gold, “IR Contaminants in Structural Firefighting,” Final Report to the National Fire Prevention and Control Administration and the Society of Plastics Industry Inc., Harvard School of Public Health, Cambridge, MA (1979).

    Google Scholar 

  109. D.A. Purser, “The Development of Toxic Hazard in Fires from Polyurethane Foams and the Effects of Fire Retardants,” in Proceedings of Flame Retardants 90, (British Plastics Federation, ed.), Elsevier, London, pp. 206–221 (1990).

    Google Scholar 

  110. R.C. Anderson and Y.C. Alarie, “Screening Procedures to Recognize ‘Supertoxic’ Decomposition Products from Polymeric Materials Under Thermal Stress,” Journal of Combustion Toxicology, 5, pp. 54–63 (1978).

    Google Scholar 

  111. W.E. Coleman, L.D. Scheel, R.E. Kupel, and R.L. Larkin, “The Identification of Toxic Compounds in the Pyrolysis Products of Polytetrafluoroethylene (PTFE),” American Industrial Hygiene Association Journal, 29, pp. 33–40 (1968).

    Article  Google Scholar 

  112. S.J. Williams and F.B. Clarke, “Combustion Product Toxicity: Dependence on the Mode of Product Generation,” Fire Materials, 6, pp. 161–162 (1982).

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Editor information

Editors and Affiliations

Appendix

Appendix

Acidosis:

A condition in which the pH of the blood is lowered (i.e., becomes more acidic). Respiratory acidosis in fire exposures results from excess carbon dioxide uptake. Metabolic acidosis results from impaired tissue respiration (due to tissue hypoxia) caused by burns or asphyxia. (See alkalosis.)

Addition:

Two or more toxic substances are considered to exert an additive effect when they act in concert, such that the effect in combination is greater than the effect of either substance acting alone but not greater than the sum of the effects of either substance acting alone (when they may be said to be directly additive). (See also synergism.)

Aerodynamic diameter:

The aerodynamic diameter of a particle is an expression of particle size, and represents the diameter of a spherical particle of unit density with the same aerodynamic properties as the particle under consideration.

Aerosol:

Solid or liquid particles dispersed in air.

Alkalosis:

Respiratory alkalosis occurs when the pH of the blood is increased (i.e., becomes more alkaline). It is caused by excess removal of carbon dioxide from the blood via the lungs during hyperventilation and may cause a loss of consciousness.

Available Safe Escape Time (ASET):

ISO Definition: Time available for escape for an individual occupant, the calculated time interval between the time of ignition and the time at which conditions become such that the occupant is estimated to be incapacitated, i.e., unable to take effective action to escape to a safe refuge or place of safety

NOTE 1 The time of ignition can be known, e.g., in the case of a fire model or a fire test or it may be assumed, e.g., it may be based upon an estimate working back from the time of detection. The basis on which the time of ignition is determined is always stated.

NOTE 2 This definition equates incapacitation with failure to escape. Other criteria for ASET are possible. If an alternate criterion is selected, it is necessary that it be stated.

NOTE 3 Each occupant can have a different value of ASET, depending on that occupant’s personal characteristics.

Asphyxia:

Suffocation; a decrease in the oxygen content, and increase in the carbon dioxide content of the blood that may occur due to laryngeal spasm caused by burns or irritant gases or to impairment of breathing or gas exchange in the lung. The term has been extended to include all causes of tissue hypoxia, including exposure to asphyxiant gases (low oxygen concentration due to the excess of any other gas, or exposure to the asphyxiant gases carbon monoxide and hydrogen cyanide, which produce asphyxia chemically).

Atmosphere (fire atmosphere or test atmosphere):

The total airborne medium to which a victim or experimental animal is exposed, consisting of solid and liquid particles and vapors dispersed in air.

Behavioral effects/incapacitation:

The extent to which exposure to fire products affects the ability or willingness of a subject or experimental animal to perform coordinated movements or tasks, particularly movements or tasks similar to those required to escape from a fire. (See incapacitation.)

Bioassay:

Originally a term reserved for the use of a biological system to detect or measure the amount of a biologically active material. In the fire context, it refers to the use of animal exposures rather than chemical analysis to determine the toxicity of a combustion product atmosphere.

Blepharospasm:

Involuntary and sustained closure (spasm) of the eyelids. In fires it is due to the painful stimulation of the cornea by combustion products that are sensory irritants.

Bronchoconstriction:

Constriction of the conducting airways in the lung due to the contraction of smooth muscle in the airway walls in response to an agonist or to stimulation of irritant receptors acting through the vagus nerve.

Burn:

Tissue lesion caused by heat or chemicals. For description of burn types and degrees, see text.

Carboxyhemoglobin (COHb):

Combination of carbon monoxide with hemoglobin in the blood, which limits the combination of hemoglobin with oxygen (oxyhemoglobin), and therefore the carriage of oxygen in the blood.

Cerebral depression:

Condition in which the electrical activity of the cerebral cortex as revealed in the electroencephalogram consists mainly of slow wave (or delta wave) activity that is typical of a semiconscious or unconscious state.

Combustion products:

Strictly speaking, this term means the products of flaming decomposition and is used in this sense when contrasted with thermal decomposition products. However, in general usage, the term may be taken to include all fire products, whether produced by flaming or nonflaming thermal decomposition.

Concentration:

The amount of a contaminant in the atmosphere per unit volume of the atmosphere, usually quoted as mass/volume (mg/L or mg/m3) or volume/ volume (ppm or percent). (See nominal atmosphere concentration.)

Dose:

The amount of a toxicant to which a fire victim or test animal is exposed. The simplest estimation of dose for inhalation toxicology is to multiply the atmosphere concentration by the duration of exposure (Ct product). A lethal dose may be expressed in terms of the LCt 50. However, other factors may affect the amount of toxicant actually entering the body, and for fires it may be necessary to express dose in terms of the material in the fire. (See nominal atmosphere concentration.)

Edema:

Accumulation of an excessive amount of fluid in cells, tissues, or body cavities. Pulmonary edema occurs when a fluid exudate leaks out of blood vessels as a result of inflammation or circulatory insufficiency, and the lung tissues become swollen and waterlogged. Further development results in a fluid exuded within the alveolar spaces. This fluid accumulation seriously affects gas exchange in the lung and may be fatal.

Electroencephalogram:

Waves of electrical activity in the cerebral cortex recorded from the surface of the head, which give an indication of the physiological state of the brain and the degree of alertness of the subject. A preponderance of fast (beta and alpha) activity indicates a conscious and normal state, whereas a preponderance of slow (theta and delta) activity signifies a physiologically depressed or unconscious state.

Erythema:

Reddening of the skin in response to heat. This change coincides with pain and just precedes a skin burn.

Fire profile:

Record of the changes with time of the concentrations of important fire products and intensities of physical parameters during the course of a fire.

Flaming fire:

In the context of this chapter, this term refers to the early stages of fire growth (preflashover), when the fire is still confined to burning items within a well-defined area.

Flashover:

Point in growth of a flaming fire where the flames are no longer confined to burning items but also occur within the fire effluent, remote from the seat of the fire.

Fractional incapacitating dose:

The dose of a toxic product acquired during a short period of time, expressed as a fraction of the dose required to cause incapacitation at the average exposure concentration during that time interval. The fractional incapacitating doses acquired during each short time period are summed throughout the exposure, incapacitation occurring when the fraction reaches unity.

Fully developed fire:

A fire that has reached its maximum extent of growth, usually extending throughout the fire compartment.

Haber’s rule:

Principle that toxicity in inhalation toxicology depends on the dose available and that the product of concentration and exposure time is a constant.

Hazard:

A toxic fire hazard exists when a toxic product is present at a sufficient concentration and over a sufficient period of time to cause a toxic effect. A physical fire hazard exists when a physical fire parameter (heat or smoke) is present at an intensity and over a period sufficient to cause injury or seriously inhibit the ability to escape from a fire.

Hypercapnia:

Increased blood carbon dioxide concentration.

Hyperthermia (heat stroke):

An increase in body temperature above 37 °C. Hyperthermia is life-threatening if the body core temperature, or temperature of the blood entering the heart, exceeds 42.5 °C.

Hyperventilation:

Increased rate and depth of breathing (increased respiratory minute volume, or RMV), in response to increased carbon dioxide, hypoxic hypoxia, hydrogen cyanide, exercise, heat, or stimulation of pulmonary irritant receptors.

Hypoxia:

A reduction in the amount of oxygen available for tissue respiration, which can occur in the following four ways:

Anemic hypoxia:

The arterial PO2 is normal, but the amount of hemoglobin available to carry oxygen is reduced and the ability to release oxygen to the tissues is impaired. For fire exposures this results mainly from the formation of carboxyhemoglobin following exposure to CO, but an anemic subject would be at increased risk.

Histotoxic hypoxia:

The amount of oxygen delivered to the tissues is adequate, but due to the action of a toxic agent such as HCN, the tissue cells cannot make use of the oxygen supplied to them.

Hypoxic hypoxia (low-oxygen hypoxia):

The PO2 of the arterial blood is reduced as a result of a low atmospheric oxygen concentration or impairment of gas exchange in the lung, due to bronchoconstriction or respiratory tract damage or disease.

Ischemic hypoxia:

Blood flow to a tissue is so low that adequate oxygen is not delivered to it despite a normal PO2 and hemoglobin concentration. This occurs during shock following burns and in cerebral tissue due to alkalosis or briefly during postural hypotension.

Incapacitation:

An inability to perform a task (related to escape from a fire) caused by exposure to a toxic substance or physical agent in a fire. A distinction is sometimes made between severe physiological incapacitation, in which the subject is unable to move normally, such as might occur in an unconscious or badly burned victim, and the more behavioral incapacitation, such as that caused by visual obscuration or eye irritation from smoke, in which the victim is more or less intact, but still unable to escape from the fire.

Inflammation:

A complex of reactions occurring in blood vessels and adjacent tissues around the site of an injury. The initial reaction is congestion (engorgement of local blood vessels), exudation of fluid into the tissues (edema), and pain followed by a phase of destruction and removal of injured tissue by inflammatory cells and then a phase of repair.

Intensity:

Level of a harmful physical fire parameter (such as radiant heat flux, air temperature, or smoke optical density).

Intoxication:

A state in which a subject is adversely affected by a toxic substance. Specifically, the time at which a subject has taken up a sufficient amount of an asphyxiant (narcotic) gas that he or she behaves like someone severely affected by alcohol.

Irritation and irritancy:

Irritation is the action of an irritant substance, irritancy is the response. This response takes the following two forms:

Pulmonary (lung) irritant:

Response occurs when an irritant penetrates into the lower respiratory tract and may result in breathing discomfort (dyspnea), bronchoconstriction, and an increase in respiratory rate during the fire exposure. In severe cases it is followed after a period (usually of a few hours) by pulmonary inflammation and edema, which may be fatal.

Sensory irritant:

Response occurs when an irritant substance comes in contact with the eyes and upper respiratory tract (and sometimes the skin), causing a painful sensation accompanied by inflammation with lacrimation or mucus secretion. At low concentrations, this effect adds to the visual obscuration caused by smoke, but at high concentrations the severe effects may cause behavioral, and to some extent physiological, incapacitation. Sensory irritation causes a decrease in respiratory rate that is transient in humans but continuous in rodents.

Lacrimation:

The production of tears in response to sensory irritation of the eyes.

LC 50 :

Lethal concentration—50 %. The concentration statistically calculated to cause the deaths of one-half of the animals exposed to a toxicant for a specified time. It may be expressed as volume/volume (ppm, percent) or mass/volume (mg/L). Care must be taken in comparing LC 50s of both the exposure duration and the postexposure period over which deaths were scored. In combustion toxicology, the LC 50 may be related to the test material rather than its products and expressed in terms of the nominal atmosphere concentration of material either of mass charge or mass loss. (See nominal atmosphere concentration.)

LCt 50 :

The product of exposure concentration and duration causing the deaths of 50 % of animals.

Narcosis:

Literally “sleep induction” but used in combustion toxicology to describe central nervous system depression causing reduced awareness, intoxication, and reduced escape capability, leading to loss of consciousness and death in extreme cases. The asphyxiant gases CO, HCN, and CO2 cause asphyxia, as does lack of oxygen due to the inhalation of an atmosphere low in oxygen, an impairment of breathing, or an impairment of gas exchange in the lung. The terms narcosis and narcotic gases are used synonymously with the terms asphxia and asphyxiant gases.

Nominal Atmosphere Concentration (NAC):

The theoretical concentration of test substance in a test atmosphere, calculated from the mass of test substance produced from the atmosphere generation system each minute divided by the air volume into which it is generated. This concept is not directly applicable to combustion toxicology since the test material is decomposed in the fire or furnace system, but two derivative concepts are used to relate the test material to the degree of toxicity as follows:

Nominal Atmosphere Concentration mass charge (NAC mass charge):

The mass of material placed in the furnace system per volume of air into which it is dispersed (mg material/liter).

Nominal Atmosphere Concentration mass loss (NAC mass loss):

The mass loss of material during decomposition per volume of air into which it is dispersed (mg material/liter).

Physiological effects:

Effects of chemical fire products or physical fire parameters on the functioning of the body, as opposed to parameters affecting the mind. Thus, a physiological effect of smoke is that it obscures vision, which might have a psychological effect on the willingness of a victim to enter a smoke-filled corridor.

Pneumonia (Pneumonitis):

Inflammation of the lungs, in fire victims due to the direct effects of inhaled chemicals or hot gases, or secondarily to skin burns. The initial inflammatory phase may be followed by infection. As it passes through different phases, pneumonia may be life-threatening at any time from 1 h after exposure in a fire to several weeks after exposure.

Potency:

The toxic potency is a measure of the amount of a toxic substance required to elicit a specific toxic effect—the smaller the amount required, the greater the potency.

Psychological effects:

Psychological effects of exposure to fire scenarios are on the mind of the victim and may result in a variety of behavioral effects. These are distinct from physiological effects on body function (see above). A fire victim is likely to suffer both types of effects at various stages of a fire, and interactions between psychological and physiological effects are likely.

Psychomotor:

Psychomotor skills are required to perform behavioral tasks involving a series of coordinated movements of the type required to escape from a fire in a compartment (such as a building).

Pyrolysis:

In this chapter, the term pyrolysis is restricted to the thermal decomposition of materials without oxidation. In small-scale tests pyrolysis may be achieved by heating the material in a stream of nitrogen.

RD 50 :

Respiratory depression 50 %—statistically calculated concentration of a sensory irritant required to reduce the breathing rate of laboratory rodents (usually mice) by 50%.

Respiratory Minute Volume (RMV):

Volume of air breathed each minute (liters/minute). RMV = TV × RR.

Respiratory Rate (RR):

Respiratory frequency (i.e., number of breaths per minute).

Respiratory tract:

The nose, pharynx, larynx, trachea, and large bronchi are termed the upper respiratory tract, and the bronchiole, alveolar ducts, and alveoli are termed the lower respiratory tract.

Required Safe Escape Time (RSET) ISO definition:

Time required for escape. Calculated time period required for an individual occupant to travel from their location at the time of ignition to a safe refuge or place of safety cf. available safe escape time and evacuation time.

Shock:

A reduction in the circulating blood volume with a fall in blood pressure.

Smoke:

Total fire effluents, consisting of solid and liquid particles and vapors.

Smoldering/nonflaming oxidative decomposition:

Thermal decomposition in which there is partial oxidation of the pyrolysis products but no flame. This may result from overheating of materials by means of an external heat source or from self-sustained smoldering.

Specific toxicity:

A particular adverse effect caused by a toxicant (e.g., asphyxia, irritancy).

Supertoxicant:

A term used to describe a toxicant with an unusual specific toxicity not usually associated with fire effluents, often with a high potency.

Synergism:

Situation where the toxic potency of two or more substances acting in concert is greater than the sum of the potencies of each substance acting alone.

Tenability limit:

Maximum concentration of a toxic fire product or intensity of a physical fire parameter that can be tolerated without causing incapacitation.

Thermal decomposition:

Chemical breakdown of a material induced by the application of heat.

Tidal Volume (TV):

Volume of air exhaled in each breath.

Toxicity:

The nature and extent of adverse effects of a substance on a living organism.

Ventilation (lung):

The volume of air breathed each minute (synonymous with respiratory minute volume).

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Society of Fire Protection Engineers

About this chapter

Cite this chapter

Purser, D.A. (2016). Combustion Toxicity. In: Hurley, M.J., et al. SFPE Handbook of Fire Protection Engineering. Springer, New York, NY. https://doi.org/10.1007/978-1-4939-2565-0_62

Download citation

  • DOI: https://doi.org/10.1007/978-1-4939-2565-0_62

  • Publisher Name: Springer, New York, NY

  • Print ISBN: 978-1-4939-2564-3

  • Online ISBN: 978-1-4939-2565-0

  • eBook Packages: EngineeringEngineering (R0)

Publish with us

Policies and ethics