Thursday, 24 March 2011 16:50

Measurement Strategies and Techniques for Occupational Exposure Assessment in Epidemiology

Rate this item
(0 votes)

Other articles in this chapter present general principles of medical surveillance of occupational illnesses and exposure surveillance. This article outlines some principles of epidemiological methods that may be used to fulfil surveillance needs. Application of these methods must take into account basic principles of physical measurement as well as standard epidemiological data-gathering practice.

Epidemiology can quantify the association between occupational and non-occupational exposure to chemico-physical stressors or behaviour and disease outcomes, and can thus provide information to develop interventions and prevention programmes (Coenen 1981; Coenen and Engels 1993). Availability of data and access to workplace and personnel records usually dictate the design of such studies. Under the most favourable circumstances, exposures can be determined through industrial hygiene measurements that are carried out in an operating shop or factory, and direct medical examinations of workers are used to ascertain possible health effects. Such evaluations can be done prospectively for a period of months or years to estimate risks of diseases such as cancer. However, it is more often the case that past exposures must be reconstructed historically, projecting backwards from current levels or using measurements recorded in the past, which may not completely meet informational needs. This article presents some guidelines and limitations for measurement strategies and documentation that affect epidemiological assessment of workplace health hazards.

Measurements

Measurements should be quantitative wherever possible, rather than qualitative, because quantitative data are subject to more powerful statistical techniques. Observable data are commonly classified as nominal, ordinal, interval and ratio. Nominal level data are qualitative descriptors which differentiate only types, such as different departments within a factory or different industries. Ordinal variables may be arranged from “low” to “high” without conveying further quantitative relationships. An example is “exposed” vs. “unexposed”, or classifying smoking history as non-smoker (= 0), light smoker (= 1), medium smoker (= 2) and heavy smoker (= 3). The higher the numerical value, the stronger the smoking intensity. Most measurement values are expressed as ratio or interval scales, in which a concentration of 30 mg/m3 is double the concentration of 15 mg/m3. Ratio variables possess an absolute zero (like age) while interval variables (like IQ) do not.

Measurement strategy

Measurement strategy takes into account information about the measurement site, the surrounding conditions (e.g., humidity, air pressure) during the measurement, the duration of the measurement and the measurement technique (Hansen and Whitehead 1988; Ott 1993).

Legal requirements often dictate measurement of eight-hour time-weighted averages (TWAs) of levels of hazardous substances. However, not all individuals work eight-hour shifts all the time, and levels of exposures may fluctuate during the shift. A value measured for one person’s job might be considered representative of an eight-hour shift value if the exposure duration is longer than six hours during the shift. As a practical criterion, a sampling duration of at least two hours should be sought. With time intervals that are too short, the sampling in one time period can show higher or lower concentrations, thereby over- or underestimating the concentration during the shift (Rappaport 1991). Therefore, it can be useful to combine several measurements or measurements over several shifts into a single time-weighted average, or to use repeated measurements with shorter sampling durations.

Measurement validity

Surveillance data must satisfy well-established criteria. The measurement technique should not influence the results during the measurement process (reactivity). Furthermore, the measurement should be objective, reliable and valid. The results should not be influenced either by the measurement technique used (execution objectivity) or by the reading or documentation by the measurement technician (assessment objectivity). The same measurement values should be obtained under the same conditions (reliability); the intended thing should be measured (validity) and interactions with other substances or exposures should not unduly influence the results.

Quality of Exposure Data

Data sources. A basic principle of epidemiology is that measurements made at the individual level are preferable to those made at the group level. Thus, the quality of epidemiolological surveillance data decreases in the following order:

    1. direct measurements taken of persons; information on exposure levels and time progression
    2. direct measurements taken of groups; information on current exposure levels for specific groups of workers (sometimes expressed as job-exposure matrices) and their variation over time
    3. measurements abstracted or reconstructed for individuals; estimation of exposure from company records, purchasing lists, descriptions of product lines, interviews with employees
    4. measurements abstracted or reconstructed for groups; historical estimation of group-based exposure indexes.

           

          In principle, the most precise determination of the exposure, using documented measurement values over time, should always be sought. Unfortunately, indirectly measured or historically reconstructed exposures are often the only data available for estimating exposure-outcome relationships, even though considerable deviations exist between measured exposures and exposure values reconstructed from company records and interviews (Ahrens et al. 1994; Burdorf 1995). The quality of the data declines in the order exposure measurement, activity-related exposure index, company information, employee interviews.

          Exposure scales. The need for quantitative monitoring data in surveillance and epidemiology goes considerably beyond the narrow legal requirements of threshold values. The goal of an epidemiological investigation is to ascertain dose-effect relation-ships, taking into account potentially confounding variables. The most precise information possible, which in general can be expressed only with a high scale level (e.g., ratio scale level), should be used. Separation into larger or smaller threshold values, or coding in fractions of threshold values (e.g., 1/10, 1/4, 1/2 threshold value) as is sometimes done, essentially relies on data measured on a statistically weaker ordinal scale.

          Documentation requirements. In addition to information on the concentrations and the material and time of measurement, external measurement conditions should be documented. This should include a description of the equipment used, measurement technique, reason for the measurement and other relevant technical details. The purpose of such documentation is to ensure uniformity of measurements over time and from one study to another, and to permit comparisons between studies.

          Exposure and health outcome data gathered for individuals are usually subject to privacy laws that vary from one country to another. Documentation of exposure and health conditions must adhere to such laws.

          Epidemiological Requirements

          Epidemiological studies strive to establish a causal link between exposure and disease. Some aspects of surveillance measurements that affect this epidemiological assessment of risk are considered in this section.

          Type of disease. A common starting point for epidemiological studies is the clinical observation of a surge in a particular disease in a company or area of activity. Hypotheses on potential biological, chemical or physical causal factors ensue. Depending on the availability of data, these factors (exposures) are studied using a retrospective or prospective design. The time between the beginning of the exposure and the onset of the disease (latency) also affects study design. The range of latency can be considerable. Infections from certain enteroviruses have latency/incubation times of 2 to 3 hours, whereas for cancers latencies of 20 to 30 years are typical. Therefore, exposure data for a cancer study must cover a considerably longer period of time than for an infectious disease outbreak. Exposures which began in the distant past can continue up to the onset of disease. Other diseases associated with age, such as cardiovascular disease and stroke, can appear in the exposed group after the study begins and must be treated as competing causes. It is also possible that people classified as “not sick” are merely people who have not yet manifested clinical illness. Thus, continued medical surveillance of exposed populations must be maintained.

          Statistical power. As previously stated, measurements should be expressed on as high a data level (ratio scale level) as possible in order to optimize the statistical power to produce statistically significant results. Power in turn is affected by the size of the total study population, the prevalence of exposure in that population, the background rate of illness and the magnitude of risk of the disease that is caused by the exposure under study.

          Mandated disease classification. Several systems are available for codifying medical diagnoses. The most common are ICD-9 (International Classification of Diseases) and SNOMED (Systematic Nomenclature of Medicine). ICD-O (oncology) is a particularization of the ICD for codifying cancers. ICD coding documentation is legally mandated in many health systems throughout the world, especially in Western countries. However, SNOMED codification can also codify possible causal factors and external conditions. Many countries have developed specialized coding systems to classify injuries and illnesses that also include the circumstances of the accident or exposure. (See the articles “Case study: Worker protection and statistics on accidents and occupational diseases—HVBG, Germany” and “Development and application of an occupational injury and illness classification system”, elsewhere in this chapter.)

          Measurements that are made for scientific purposes are not bound by the legal requirements that apply to mandated surveillance activities, such as determination of whether threshold limits have been exceeded in a given workplace. It is useful to examine exposure measurements and records in such a way as to check for possible excursions. (See, for example, the article “Occupational hazard surveillance” in this chapter.)

          Treatment of mixed exposures. Diseases often have several causes. Therefore it is necessary to record as completely as possible the suspected causal factors (exposures/confounding factors) in order to be able to distinguish the effects of suspected hazardous agents from one another and from the effects of other contributory or confounding factors, such as cigarette smoking. Occupational exposures are often mixed (e.g., solvent mixtures; welding fumes such as nickel and cadmium; and in mining, fine dust, quartz and radon). Additional risk factors for cancers include smoking, excess alcohol consumption, poor nutrition and age. Besides chemical exposures, exposures to physical stressors (vibration, noise, electromagnetic fields) are possible triggers for diseases and must be considered as potential causal factors in epidemiological studies.

          Exposures to multiple agents or stressors may produce interaction effects, in which the effect of one exposure is magnified or reduced by another that occurs contemporaneously. A typical example is the link between asbestos and lung cancer, which is many times more pronounced among smokers. An example of the mixture of chemical and physical exposures is progressive systemic scleroderma (PSS), which is probably caused by a combined exposure to vibration, solvent mixtures and quartz dust.

          Consideration of bias. Bias is a systematic error in classifying persons in the “exposed/not exposed” or “diseased/not diseased” groups. Two types of bias should be distinguished: observation (information) bias and selection bias. With observation (information) bias, different criteria may be used to classify subjects into the diseased/not diseased groups. It is sometimes created when the target of a study includes persons employed in occupations known to be hazardous, and who may already be under increased medical surveillance relative to a comparison population.

          In selection bias two possibilities should be distinguished. Case-control studies begin by separating persons with the disease of interest from those without that disease, then examine differences in exposure between these two groups; cohort studies determine disease rates in groups with different exposures. In either type of study, selection bias exists when information on the exposure affects classification of subjects as sick or not sick, or when information on disease status affects classification of subjects as exposed or not exposed. A common example of selection bias in cohort studies is the “healthy worker effect”, which is encountered when disease rates in exposed workers are compared with those in the general population. This can result in underestimation of disease risk because working populations are often selected from the general population on the basis of continued good health, frequently based upon medical examination, whereas the general population contains the ill and infirm.

          Confounders. Confounding is the phenomenon whereby a third variable (the confounder) alters the estimate of an association between a presumed antecedent factor and a disease. It can occur when the selection of subjects (cases and controls in a case-control study or exposed and unexposed in a cohort study) depends in some way upon the third variable, possibly in a manner unknown to the investigator. Variables associated only with exposure or disease are not confounders. To be a confounder a variable must meet three conditions:

          • It must be a risk factor for the disease.
          • It must be associated with the exposure in the study population.
          • It must not be in the causal pathway from exposure to disease.

           

          Before any data are collected for a study it is sometimes impossible to predict whether or not a variable is a likely confounder. A variable which has been treated as a confounder in a previous study might not be associated with exposure in a new study within a different population, and would therefore not be a confounder in the new study. For instance, if all subjects are alike with respect to a variable (e.g., sex), then that variable cannot be a confounder in that particular study. Confounding by a particular variable can be accounted for (“controlled”) only if the variable is measured along with exposure and illness outcomes. Statistical control of confounding may be done crudely using stratification by the con-founding variable, or more precisely using regression or other multivariate techniques.

          Summary

          The requirements of measuring strategy, measuring technology and documentation for industrial workplaces are sometimes statutorily defined in terms of threshold limit value surveillance. Data protection regulations also apply to the protection of company secrets and person-related data. These requirements call for the comparable measuring results and measurement conditions and for an objective, valid and reliable measuring technology. Additional requirements put forward by epidemiology refer to the representativeness of measurements and to the possibility of establishing links between exposures for individuals and subsequent health outcomes. Measurements may be representative for certain tasks, i.e. they may reflect typical exposure during certain activities or in specific branches or typical exposure of defined groups of persons. It would be desirable to have measurement data directly attributed to the study subjects. This would make it necessary to include with measurement documentation information about persons working at the concerned workplace during the measurement or to set up a registry allowing such direct attribution. Epidemiological data collected at the individual level are usually preferable to those obtained at the group level.

           

          Back

          Read 4941 times Last modified on Tuesday, 26 July 2022 19:24

          " DISCLAIMER: The ILO does not take responsibility for content presented on this web portal that is presented in any language other than English, which is the language used for the initial production and peer-review of original content. Certain statistics have not been updated since the production of the 4th edition of the Encyclopaedia (1998)."

          Contents

          Record Systems and Surveillance References

          Agricola, G. 1556. De Re Metallica. Translated by HC Hoover and LH Hoover. 1950. New York: Dover.

          Ahrens, W, KH Jöckel, P Brochard, U Bolm-Audorf, K Grossgarten, Y Iwatsubo, E Orlowski, H Pohlabeln, and F Berrino. 1993. Retrospective assessment of asbestos exposure. l. Case-control analysis in a study of lung cancer: Efficiency of job-specific questionnaires and job-exposure-matrices. Int J Epidemiol 1993 Suppl. 2:S83-S95.

          Alho, J, T Kauppinen, and E Sundquist. 1988. Use of exposure registration in the prevention of occupational cancer in Finland. Am J Ind Med 13:581-592.

          American National Standards Institute (ANSI). 1963. American National Standard Method of Recording Basic Facts Relating to the Nature and Occurrence of Work Injuries. New York: ANSI.

          Baker, EL. 1986. Comprehensive Plan for Surveillance of Occupational Illness and Injury in the United States. Washington, DC: NIOSH.

          Baker, EL, PA Honchar, and LJ Fine. 1989. Surveillance in occupational illness and injury: Concepts and content. Am J Public Health 79:9-11.

          Baker, EL, JM Melius, and JD Millar. 1988. Surveillance of occupational illness and injury in the United States: Current perspectives and future directions. J Publ Health Policy 9:198-221.

          Baser, ME and D Marion. 1990. A statewide case registry for surveillance of occupational heavy metals absorption. Am J Public Health 80:162-164.

          Bennett, B. 1990. World Register of Cases of Angiosarcoma of the Liver (ASL) due to Vinyl Chloride Monomer: ICI Registry.

          Brackbill, RM, TM Frazier, and S Shilling. 1988. Smoking characteristics of workers, 1978-1980. Am J Ind Med 13:4-41.

          Burdoff, A. 1995. Reducing random measurement-error in assessing postural load on the back in epidemiologic surveys. Scand J Work Environ Health 21:15-23.

          Bureau of Labor Statistics (BLS). 1986. Record Keeping Guidelines for Occupational Injuries and Illnesses. Washington, DC: US Department of Labor.

          —. 1989. California Work Injuries and Illness. Washington, DC: US Department of Labor.

          —. 1992. Occupational Injury and Illness Classification Manual. Washington, DC: US Department of Labor.

          —. 1993a. Occupational Injuries and Illnesses in the United States by Industry, 1991. Washington, DC: US Department of Labor.

          —. 1993b. Survey of Occupational Injuries and Illnesses. Washington, DC: US Department of Labor.

          —. 1994. Survey of Occupational Injuries and Illnesses, 1992. Washington, DC: US Department of Labor.

          Bureau of the Census. 1992. Alphabetic List of Industries and Occupations. Washington, DC: US Government Printing Office.

          —. 1993. Current Population Survey, January through December 1993 (Machine-Readable Data Files). Washington, DC: Bureau of the Census.

          Burstein, JM and BS Levy. 1994. The teaching of occupational health in United States medical schools. Little improvement in nine years. Am J Public Health 84:846-849.

          Castorino, J and L Rosenstock. 1992. Physician shortage in occupational and environmental medicine. Ann Intern Med 113:983-986.

          Checkoway, H, NE Pearce, and DJ Crawford-Brown. 1989. Research Methods in Occupational Epidemiology. New York: Oxford Univ. Press.

          Chowdhury, NH, C Fowler, and FJ Mycroft. 1994. Adult blood lead epidemiology and surveillance—United States, 1992-1994. Morb Mortal Weekly Rep 43:483-485.

          Coenen, W. 1981. Measurement strategies and documentation concepts for collecting hazardous work materials. Modern accident prevention (in German). Mod Unfallverhütung:52-57.

          Coenen, W and LH Engels. 1993. Mastering the risks on the job. Research for developing new preventive strategies (in German). BG 2:88-91.

          Craft, B, D Spundin, R Spirtas, and V Behrens. 1977. Draft report of a task force on occupational health surveillance. In Hazard Surveillance in Occupational Disease, edited by J Froines, DH Wegman, and E Eisen. Am J Pub Health 79 (Supplement) 1989.

          Dubrow, R, JP Sestito, NR Lalich, CA Burnett, and JA Salg. 1987. Death certificate-based occupational mortality surveillance in the United States. Am J Ind Med 11:329-342.

          Figgs, LW, M Dosemeci, and A Blair. 1995. United States non-Hodgkin’s lymphoma surveillance by occupation 1984-1989: A twenty-four-state death certificate study. Am J Ind Med 27:817-835.

          Frazier, TM, NR Lalich, and DH Pederson. 1983. Uses of computer generated maps in occupational hazard and mortality surveillance. Scand J Work Environ Health 9:148-154.

          Freund, E, PJ Seligman, TL Chorba, SK Safford, JG Drachmann, and HF Hull. 1989. Mandatory reporting of occupational diseases by clinicians. JAMA 262:3041-3044.

          Froines, JR, DH Wegman, and CA Dellenbaugh. 1986. An approach to the characterization of silica exposure in US industry. Am J Ind Med 10:345-361.

          Froines, JR, S Baron, DH Wegman, and S O’Rourke. 1990. Characterization of the airborne concentrations of lead in US industry. Am J Ind Med 18:1-17.

          Gallagher, RF, WJ Threlfall, PR Band, and JJ Spinelli. 1989. Occupational Mortality in British Columbia 1950-1984. Vancouver: Cancer Control Agency of British Columbia.

          Guralnick, L. 1962. Mortality by occupation and industry among men 20-46 years of age: United States, 1950. Vital Statistics-Special Reports 53 (2). Washington, DC: National Center for Health Statistics.

          —. 1963a. Mortality by industry and cause of death among men 20 to 40 years of age: United States, 1950. Vital Statistics-Special Reports, 53(4). Washington, DC: National Center for Health Statistics.

          —. 1963b. Mortality by occupation and cause of death among men 20 to 64 years of age: United States, 1950. Vital Statistics-Special Reports 53(3). Washington, DC: National Center for Health Statistics.

          Halperin, WE and TM Frazier. 1985. Surveillance for the effects of workplace exposure. Ann Rev Public Health 6:419-432.

          Hansen, DJ and LW Whitehead. 1988. The influence of task and location on solvent exposures in a printing plant. Am Ind Hyg Assoc J 49:259-265.

          Haerting, FH and W Hesse. 1879. Der Lungenkrebs, die Bergkrankheit in den Schneeberger Gruben Vierteljahrsschr gerichtl. Medizin und Öffentl. Gesundheitswesen 31:296-307.

          Institute of Medicine. 1988. Role of the Primary Care Physician in Occupational and Environmental Medicine. Washington, DC: National Academy Press.

          International Agency for Research on Cancer (IARC). 1990. Phenoxy acid herbicides and contaminants: Description of the IARC international register of workers. Am J Ind Med 18:39-45.

          International Labour Organization (ILO). 1980. Guidelines for the Use of ILO International Classification of Radiographs of Pneumoconioses. Occupational Safety and Health Series, No. 22. Geneva: ILO.

          Jacobi, W, K Henrichs, and D Barclay. 1992. Verursachungswahrscheinlichkeit von Lungenkrebs durch die berufliche Strahlenexposition von Uran-Bergarbeitem der Wismut AG. Neuherberg: GSF—Bericht S-14/92.

          Jacobi, W and P Roth. 1995. Risiko und Verursachungs-Wahrscheinlichkeit von extrapulmonalen Krebserkrankungen durch die berufliche Strahlenexposition von Beschäftigten der ehemaligen. Neuherberg: GSF—Bericht S-4/95.

          Kauppinen, T, M Kogevinas, E Johnson, H Becher, PA Bertazzi, HB de Mesquita, D Coggon, L Green, M Littorin, and E Lynge. 1993. Chemical exposure in manufacture of phenoxy herbicides and chlorophenols and in spraying of phenoxy herbicides. Am J Ind Med 23:903-920.

          Landrigan, PJ. 1989. Improving the surveillance of occupational disease. Am J Public Health 79:1601-1602.

          Lee, HS and WH Phoon. 1989. Occupational asthma in Singapore. J Occup Med, Singapore 1:22-27.

          Linet, MS, H Malker, and JK McLaughlin. 1988. Leukemias and occupation in Sweden. A registry-based analysis. Am J Ind Med 14:319-330.

          Lubin, JH, JD Boise, RW Hornung, C Edling, GR Howe, E Kunz, RA Kusiak, HI Morrison, EP Radford, JM Samet, M Tirmarche, A Woodward, TS Xiang, and DA Pierce. 1994. Radon and Lung Cancer Risk: A Joint Analysis of 11 Underground Miners Studies. Bethesda, MD: National Institute of Health (NIH).

          Markowitz, S. 1992. The role of surveillance in occupational health. In Environmental and Occupational Medicine, edited by W Rom.

          Markowitz, SB, E Fischer, MD Fahs, J Shapiro, and P Landrigan. 1989. Occupational disease in New York State. Am J Ind Med 16:417-435.

          Matte, TD, RE Hoffman, KD Rosenman, and M Stanbury. 1990. Surveillance of occupational asthma under the SENSOR model. Chest 98:173S-178S.

          McDowell, ME. 1983. Leukemia mortality in electrical workers in England and Wales. Lancet 1:246.

          Melius, JM, JP Sestito, and PJ Seligman. 1989. Occupational disease surveillance with existing data sources. Am J Public Health 79:46-52.

          Milham, S. 1982. Mortality from leukemia in workers exposed to electrical and magnetic fields. New Engl J Med 307:249.

          —. 1983. Occupational Mortality in Washington State 1950-1979. NIOSH publication No. 83-116. Springfield, Va: National Technical Information Service.

          Muldoon, JT, LA Wintermeyer, JA Eure, L Fuortes, JA Merchant, LSF Van, and TB Richards. 1987. Occupational disease surveillance data sources 1985. Am J Public Health 77:1006-1008.

          National Research Council (NRC). 1984. Toxicity Testing Strategies to Determine Needs and Priorities. Washington, DC: National Academic Press.

          Office of Management and Budget (OMB). 1987. Standard Industrial Classification Manual. Washington, DC: US Government Printing Office.

          OSHA. 1970. The Occupational Safety and Health Act of 1970 Public Law 91-596 91st US Congress.

          Ott, G. 1993. Strategic proposals for measurement technique in occurrences of damage (in German). Dräger Heft 355:2-5.

          Pearce, NE, RA Sheppard, JK Howard, J Fraser, and BM Lilley. 1985. Leukemia in electrical workers in New Zealand. Lancet ii:811-812.

          Phoon, WH. 1989. Occupational diseases in Singapore. J Occup Med, Singapore 1:17-21.

          Pollack, ES and DG Keimig (eds.). 1987. Counting Injuries and Illnesses in the Workplace: Proposals for a Better System. Washington, DC: National Academy Press.

          Rajewsky, B. 1939. Bericht über die Schneeberger Untersuchungen. Zeitschrift für Krebsforschung 49:315-340.

          Rappaport, SM. 1991. Assessment of long-term exposures to toxic substances in air. Ann Occup Hyg 35:61-121.

          Registrar General. 1986. Occupation Mortality, Decennial Supplement for England and Wales, 1979-1980, 1982-1983 Part I Commentary. Series DS, No. 6. London: Her Majesty’s Stationery Office.

          Robinson, C, F Stern, W Halperin, H Venable, M Petersen, T Frazier, C Burnett, N Lalich, J Salg, and J Sestito. 1995. Assessment of mortality in the construction industry in the United States, 1984-1986. Am J Ind Med 28:49-70.

          Roche, LM. 1993. Use of employer illness reports for occupational disease surveillance among public employees in New Jersey. J Occup Med 35:581-586.

          Rosenman, KD. 1988. Use of hospital discharge data in the surveillance of occupational disease. Am J Ind Med 13:281-289.

          Rosenstock, L. 1981. Occupational medicine: Too long neglected. Ann Intern Med 95:994.

          Rothman, KJ. 1986. Modern Epidemiology. Boston: Little, Brown & Co.

          Seifert, B. 1987. Measurement strategy and measurement procedure for investigations of inside air. Measurement technique and Environmental protection (in German). 2:M61-M65.

          Selikoff, IJ. 1982. Disability Compensation for Asbestos-Associated Disease in the United States. New York: Mt. Sinai School of Medicine.

          Selikoff, IJ, EC Hammond, and H Seidman. 1979. Mortality experience of insulation workers in the United States and Canada, 1943-1976. Ann NY Acad Sci 330:91-116.

          Selikoff, IJ and H Seidman. 1991. Asbestos-associated deaths among insulation workers in the United States and Canada, 1967-1987. Ann NY Acad Sci 643:1-14.

          Seta, JA and DS Sundin. 1984. Trends of a decade—A perspective on occupational hazard surveillance 1970-1983. Morb Mortal Weekly Rep 34(2):15SS-24SS.

          Shilling, S and RM Brackbill. 1987. Occupational health and safety risks and potential health consequences perceived by US workers. Publ Health Rep 102:36-46.

          Slighter, R. 1994. Personal communication, United States Office of Worker’s Compensation Program, September 13, 1994.

          Tanaka, S, DK Wild, PJ Seligman, WE Halperin, VJ Behrens, and V Putz-Anderson. 1995. Prevalence and work-relatedness of self-reported carpal tunnel syndrome among US workers—Analysis of the occupational health supplement data of 1988 national health interview survey. Am J Ind Med 27:451-470.

          Teschke, K, SA Marion, A Jin, RA Fenske, and C van Netten. 1994. Strategies for determining occupational exposure in risk assessment. A review and a proposal for assessing fungicide exposures in the lumber industry. Am Ind Hyg Assoc J 55:443-449.

          Ullrich, D. 1995. Methods for determining indoor air pollution. Indoor air quality (in German). BIA-Report 2/95,91-96.

          US Department of Health and Human Services (USDHHS). 1980. Industrial Characteristics of Persons Reporting Morbidity During the Health Interview Surveys Conducted in 1969-1974. Washington, DC: USDHHS.

          —. July 1993. Vital and Health Statistics Health Conditions among the Currently Employed: United States 1988. Washington, DC: USDHHS.

          —. July 1994. Vital and Health Statistics Plan and Operation of the Third National Health and Nutrition Examination Survey, 1988-94. Vol. No. 32. Washington, DC: USDHHS.

          US Department of Labor (USDOL). 1980. An Interim Report to Congress on Occupational Diseases. Washington, DC: US Government Printing Office.

          US Public Health Services (USPHS). 1989. The International Classification of Diseases, 9th Revision, Clinical Modification. Washington, DC: US Government Printing Office.

          Wegman, DH. 1992. Hazard surveillance. Chap. 6 in Public Health Surveillance, edited by W Halperin, EL Baker, and RR Ronson. New York: Van Nostrand Reinhold.

          Wegman, DH and JR Froines. 1985. Surveillance needs for occupational health. Am J Public Health 75:1259-1261.

          Welch, L. 1989. The role of occupational health clinics in surveillance of occupational disease. Am J Public Health 79:58-60.

          Wichmann, HE, I Brüske-Hohlfeld, and M Mohner. 1995. Stichprobenerhebung und Auswertung von Personaldaten der Wismut Hauptverband der gewerblichen Berufsgenossenschaften. Forschungsbericht 617.0-WI-02, Sankt Augustin.

          World Health Organization (WHO). 1977. Manual of the International Statistical Classification of Diseases, Injuries, and Causes of Death, Based on the Recommendations of the Ninth Revision Conference, 1975. Geneva: WHO.