National Academies Press: OpenBook

Gulf War and Health: Volume 2: Insecticides and Solvents (2003)

Chapter: 6. Cancer and Exposure to Solvents

« Previous: 5. Cancer and Exposure to Insecticides
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

6
CANCER AND EXPOSURE TO SOLVENTS

INTRODUCTION

The associations between exposure to organic solvents and the incidence of and mortality from cancer have been investigated extensively in a number of industries and occupations, including dry-cleaning, painting, printing, and rubber and shoe manufacturing. As a result, the body of evidence on exposure to organic solvents and cancer reviewed by the committee is quite large compared with that on other health effects. To help the reader become more familiar with the studies on exposure to solvents and cancer, this chapter is organized differently from Chapter 5.

Following this general introduction is a description of the major occupational cohort studies that are cited throughout the chapter; these studies provide findings for multiple cancer outcomes. The committee describes the essential study design characteristics and pertinent information for each of these cohorts organized by the type of solvent exposure.

The occupational cohort studies examined populations with known or suspected exposure to the solvents under review. Many of them have been updated and expanded to include more cohort members, longer periods of assessment, and other estimates of exposure. All of the various studies that follow a particular cohort, such as the NIOSH Pliofilm cohort, are described together in Table 6.1. The committee reviewed all the papers related to each major cohort in drawing its conclusions of association, but it is usually the findings from the most recent papers that are provided in the data-analysis tables at the end of the section on each cancer site. In some cases, results from the earlier papers were never reproduced, so the committee used the earlier results in its analysis.

A description of key case-control studies at the beginning of each section is followed by a table that outlines the studies’ characteristics and design elements and is similar to the table of cohort studies at the beginning of the chapter. Discussions of strengths and limitations of the studies that formed the basis of the committee’s conclusions are presented in the sections on the specific outcomes.

Background information on the types of cancer or cancer in general is provided in Chapter 5, and the reader is referred to those sections for that information. A review of the pertinent toxicology and findings from other organizations charged with evaluating the carcinogenicity of organic solvents is provided at the end of this introduction.

As in Chapter 5, the cancer outcomes are presented in the order of the ninth revision of the International Classification of Disease (ICD-9).

The Literature on Exposure to Organic Solvents

The literature on exposure to organic solvents and cancer outcomes provides information on specific solvent exposures (for example, benzene, trichloroethylene, and

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

tetrachloroethylene), on mixtures of specific organic solvents, or on mixtures of unspecified solvents. In many studies, exposure to solvents was not assessed specifically; rather, surrogates of exposure were used, such as job title, industry type, or occupation.

As is discussed in Chapter 2, a study’s ability to determine exposure accurately and specifically is critical in evaluating its overall quality. For the purposes of this report, the committee used studies that assessed exposure to specific organic solvents or to solvent mixtures as the primary evidence for its conclusions. The committee also included surrogates of exposure in drawing its conclusions, but it viewed those studies as supportive or supplemental evidence. Those studies included exposure of painters, printers, aircraft maintenance and manufacturing workers, service-station attendants, and shoe and boot manufacturers. All those studies are included in the data-analysis tables that accompany discussions of each cancer outcome.

The committee found most of the cancer literature focused on the following compounds: benzene, trichloroethylene, tetrachloroethylene, methylene chloride, and mixtures of unspecified organic solvents. Therefore, most of the committee’s conclusions on cancer outcomes are focused on exposure to those compounds. A smaller number of studies analyzed associations between cancer outcomes and toluene, xylene, isopropyl alcohol, methyl ethyl ketone, phenol, and other individual solvents, but for most agents, there was insufficient evidence for the committee to draw conclusions.

For exposure to tetrachloroethylene, the committee included studies of dry-cleaning and laundry workers as part of the primary evidence in drawing conclusions of associations. As a result, the conclusions related to exposure to tetrachloroethylene are also related to exposure to “dry-cleaning solvents.” The committee acknowledges that dry cleaners and launderers are likely exposed to other organic solvents and chemicals, including naphtha, Stoddard solvent, carbon tetrachloride, trichloroethylene, and 1,1,1-trichloroethane (IARC, 1995). As a result, the committee decided to consider studies of both tetrachloroethylene and dry-cleaning solvents in forming their conclusions, thereby including the possibility that one of the other solvents used in that industry contributed to the risks observed in some of the studies on dry cleaners and launderers.

The committee based its review of cancer outcomes only on studies of humans that had a comparison or control group (cohort and case-control studies). Case reports, case series, review articles, and meta-analyses related to cancer were excluded from the committee’s review. Although the committee reviewed ecologic, cross-sectional, proportionate mortality ratio (PMR), and mortality odds ratio studies, it did not consider them critical to its decision and excluded them from the discussions. Chapter 2 describes their specific limitations.

Toxicity and Determination of Carcinogenicity

Excess incidence of cancer has been observed in animals exposed to the specific organic solvents reviewed by the committee. Benzene, perhaps the most thoroughly investigated solvent, is a well-established carcinogen and has repeatedly been shown to induce hematopoietic cancers and cancers of the ovaries, mammary glands, pancreas, and liver (ATSDR, 1997a). The International Agency for Research on Cancer (IARC) has determined that benzene is “carcinogenic to humans” as determined in studies of both humans and animals. IARC bases its determination of benzene’s carcinogenicity on evidence from human studies that is considered “sufficient,” whereas the available animal

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

data are considered “limited.” Most of the human studies cited by IARC involve the increased risk of leukemia and other lymphatic and hematopoietic cancers (IARC, 1987). The National Toxicology Program (NTP) has also classified benzene as “known to be a human carcinogen” in its most recent report on carcinogens on the basis of animal and human studies (NTP, 2001).

On the basis of animal studies, trichloroethylene has been associated with liver cancer in one strain of one species (mouse) (ATSDR, 1997b). Liver and renal cell cancers and mononuclear cell leukemia have typically been seen after exposure to tetrachloroethylene. According to the Agency for Toxic Substance and Disease Registry (ATSDR, 1997c), the relevance to humans of rodent toxicology studies on trichloroethylene and tetrachloroethylene is unclear, given that some mechanisms of action differ. However, a great deal of research has been conducted over the last decade, and some mechanisms of action appear to be similar in rodents and humans such as genotoxic and cytotoxic actions of mercapturic acid derivatives of both trichloroethylene and tetrachloroethylene in the kidney (see Chapter 4 for more information). IARC has also reviewed trichloroethylene and tetrachloroethylene and determined that both are “probably carcinogenic to humans.” The evidence from animal studies is stronger and considered to be “sufficient,” whereas the evidence from human studies is considered “limited” (IARC, 1995). In addition, the NTP has identified both trichloroethylene and tetrachloroethylene as “reasonably anticipated to be human carcinogens” (NTP, 2001).

Exposure to methylene chloride in some rodent species has consistently produced excess numbers of cancers of the liver and lung and benign mammary tumors (ATSDR, 2000). IARC has determined that exposure to methylene chloride is “possibly carcinogenic to humans,” and the NTP concluded that it is “reasonably anticipated to be a human carcinogen” (IARC, 1999; NTP, 2001). IARC has determined that toluene and xylene are “not classifiable as to their carcinogenicity to humans” in that there was inadequate evidence from studies of humans and animals (IARC, 1999).

Chloroform has produced liver and kidney tumors in a strain-, sex-, species-, and dose-dependent manner and, on the basis of sufficient evidence from animal studies, is “reasonably anticipated to be a human carcinogen” according to the NTP (ATSDR, 1997d; NTP, 2001). Chloroform was once used as an anesthetic, but its association with cancer in nonmedical exposures in humans has not been investigated extensively. The committee did not review studies on the efficacy of solvents as therapeutic agents (see Chapter 2). Chapter 4 provides details on the adverse effects of chloroform as observed in experimental studies.

In addition to evaluating the carcinogenicity of specific chemical agents, IARC has analyzed whether particular occupations pose a greater risk for exposure to carcinogenic agents. In fact, IARC has determined that working in the rubber industry and in boot and shoe manufacturing and repair pose such a risk (IARC, 1987), and it determined that there is “sufficient evidence for the carcinogenicity of occupational exposure as a painter” (IARC, 1989). Although IARC identifies exposures of concerns and specific cancer outcomes that demonstrate an increased risk, its overall charge is to determine whether a specific agent or occupation is carcinogenic, not whether an agent causes a specific cancer outcome. It is important to distinguish the objectives of IARC’s program and the charge of the present committee. The purpose of the IARC program is to determine whether agents or occupational exposures are carcinogenic, whereas this committee is charged with determining whether there is an association between exposure to a specific agent and

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

chronic human illnesses. As discussed earlier in this report, the committee uses experimental evidence only when it is required by the definitions of the categories of association. Only the category of “Sufficient Evidence of a Causal Association” requires support from experimental evidence. For each conclusion of causality, the animal data that provides a plausible mechanism for the outcome being discussed are described—as in the section on chronic exposure to benzene and acute leukemia. Additional information on the toxicology and available experimental data on a number of solvents reviewed in this report can be found in Chapter 4.

DESCRIPTION OF THE COHORT STUDIES

In reviewing the published epidemiologic literature on exposure to organic solvents, the committee examined a number of occupational cohort studies that provided information on the association between cancer mortality or incidence and exposure to specific organic solvents or to mixtures of organic solvents. Deaths or incident cases of cancer in the cohort studies were recorded, exposed populations were followed over time, and the relationship between rates of cancer and exposure was assessed with statistical methods. Because the cohort studies played an important role in the committee’s conclusions and are referred to throughout this chapter, they are described here according to the solvents they examined. Table 6.1 presents for each study a description of the population, the followup period, the number of subjects, the relevant exposures, the methods used to assess exposure to organic solvents, the statistical methods, and the adjustment for potential confounding variables. Similar tables for case-control studies are found in the sections on each type of cancer.

Studies of Workers Exposed to Benzene

Benzene is used in chemical processes often as an intermediate in the manufacture of other chemicals and end products. Occupational exposure to benzene has been studied primarily in industrial workers, including rubber, chemical, and petroleum and gasoline workers. On the basis of human studies of those occupational groups and animal studies over the last 60 years, the allowable occupational health standard for benzene has steadily decreased in the United States. In 1987, the permissible exposure limit (PEL) set by the Occupational Safety and Health Administration (OSHA) was reduced from 10 parts per million (ppm) to a time weighted average over 8 hours (8-hr TWA) of 1 ppm (NIOSH, 1997).

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.1 Description of Cohort Studies Related to Exposure to Organic Solvents

Reference

Description

Study Group (N)

Comparison Group (N)

Exposure Assessment and Other Relevant Exposures

Analysis and Adjustment for Potential Confounders

Benzene

NIOSH Pliofilm Cohort

Infante et al., 1977

Mortality experience (1940–1975) of white male Pliofilm workers (at least 1 day in 1940–1949) at three Goodyear facilities in Ohio

748

(1) US white male population

Employment in a benzene-exposed occupation as verified through historical air exposure measurements

SMR

Age, time period

(2) 1447 white, male fibrous-glass construction workers in Ohio

Rinsky et al., 1981

Mortality experience (1940–1975) of

 

US white male population

Employment in a benzene-exposed occupation as verified through historical air exposure measurements

SMR

Age, sex, time period

(1) the original cohort and

(1) 748

(2) a second group of white male Pliofilm workers (at least 1 day in 1950–1959)

(2) 258

at three Goodyear facilities in Ohio

Rinsky et al., 1987

Mortality experience (1940–1981) of white male Pliofilm workers (at least 1 day in 1940–1965) at three Goodyear facilities in Ohio

1165

US white male population

Employment in a benzene-exposed occupation as verified through historical air exposure measurements, with cumulative exposure indexes

SMR

Age, time period

Paxton et al., 1994a, 1996

Mortality experience (1940–1987) of white male Pliofilm workers (at least 1 day in 1940–1965) at three Goodyear facilities in Ohio

1212

US white male population

Employment in a benzene-exposed occupation as verified through modified historical air exposure measurements, with cumulative exposure indexes

SMR

Age, time period

Paxton et al., 1994b

Mortality experience (1940–1987) of white male Pliofilm workers (at least 1 day in 1940–1965) at three Goodyear facilities in Ohio

1868

US white male population

Employment in a benzene-exposed occupation as verified through modified historical air exposure measurements, with cumulative exposure indexes

SMR, Cox proportional hazards model

Age, sex, location, time of first Pliofilm employment

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description

Study Group (N)

Comparison Group (N)

Exposure Assessment and Other Relevant Exposures

Analysis and Adjustment for Potential Confounders

Crump, 1994, 1996

Mortality experience (1940–1987) of white male Pliofilm workers (at least 1 day in 1940–1965) at three Goodyear facilities in Ohio

1717

US white male population

Employment in a benzene-exposed occupation as verified through modified historical air exposure measurements, with cumulative exposure indexes

Life-table analysis

Age, sex

Wong, 1995

Mortality experience (1940–1987) of white male Pliofilm workers (at least 1 day in 1940–1965) at three Goodyear facilities in Ohio

1868

US general population

Employment in a benzene-exposed occupation as verified through historical air exposure measurements, with cumulative exposure indexes

SMR

Age

Chinese Workers Cohort

Yin et al., 1987

Mortality experience (1972–1981) of benzene-exposed workers (at least 6 months) in China

28,460 total

15,643 men

12,817 women

28,257 unexposed

Employment in a benzene-exposed occupation as verified through historical air exposure measurements from factory records

RR, SMR

Age, sex

Yin et al., 1989

Mortality experience (1972–1981) of benzene-exposed workers (at least 6 months) in China

28,460 total

15,643 men

12,817 women

28,257 unexposed

Employment in a benzene-exposed occupation as verified through historical air exposure measurements from factory records

RR, SMR

Age, sex, smoking

Yin et al., 1994

Incidence and mortality experience (1972–1981) of benzene-exposed workers (at least 6 months) in China

28,460 total

15,643 men

12,817 women

28,257 unexposed

Employment in a benzene-exposed occupation as verified through historical air exposure measurements from factory records

SMR, Poisson

Age, sex, time of first employment

Li et al., 1994

Incidence and mortality experience (1972–1987) of benzene-exposed workers (at least 1 day) in China

74,828 total

38,833 men

35,995 women

35,805 unexposed

Employment in a benzene-exposed occupation

RR

Sex

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description

Study Group (N)

Comparison Group (N)

Exposure Assessment and Other Relevant Exposures

Analysis and Adjustment for Potential Confounders

Yin et al., 1996a,b

Incidence and mortality experience (1972–1987) of benzene-exposed workers (at least 1 day) in China

74,828 total

38,833 men

35,995 women

35,805 unexposed

Employment in a benzene-exposed occupation as verified through historical air exposure measurements from factory records

RR

Age, sex

Hayes et al., 1996

Mortality experience (1972–1987) of benzene-exposed workers (at least 1 day) in China

74,828 total

38,833 men

35,995 women

35,805 unexposed

Employment in a benzene-exposed occupation with cumulative exposure assigned by industrial hygienist from historical records

RR (Poisson), trend analysis

Age, sex

Hayes et al., 1997

Incidence (1972–1987) in benzene-exposed workers (at least 1 day) in China

74,828 total

38,833 men

35,995 women

35,805 unexposed

Employment in a benzene-exposed occupation with cumulative and average exposure assigned by industrial hygienist from historical records

RR (Poisson)

Age, sex

Other Cohort Studies

McMichael et al., 1976

Mortality experience (1964–1973) of male rubber workers (at least 1 day) in four plants in Ohio and Wisconsin

18,903

1968 US male population

Employment at one of four rubber-manufacturing plants

SMR

Age, race

Wilcosky et al., 1984

Cases, age 40–84 years, selected retrospectively from a cohort of active and retired male rubber workers in a plant in Akron, Ohio, in 1964–1973; an age-stratified, 20% random sample from the original cohort served as the control group

NA

1336 (20% of 6678)

Linkage of worker histories to plant solvent-use records; work in process area with known solvent use equates to exposure

Race-specific

ORs

Age

 

Other exposures: trichloroethylene, tetrachloroethylene, toluene, xylenes, naphthas, ethanol, acetone, phenol

 

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description

Study Group (N)

Comparison Group (N)

Exposure Assessment and Other Relevant Exposures

Analysis and Adjustment for Potential Confounders

Pippard and Acheson, 1985

Mortality experience (1939–1982) of male boot and shoe manufacturers (in 1939) in three tow ns in Great Britain

5017

County general populations

Job title

SMR

Age, time period

 

Other exposures: trichloroethylene, solvents

Wong, 1987a

Mortality experience (1946–1977) of male chemical workers (at least 6 months) in seven US plants

7676

US general population

Job title and employment duration

SMR, Mantel-Haenszel RR

Age, race

Wong, 1987b

Mortality experience (1946–1977) of male chemical workers (at least 6 months) in seven US plants

7676

US general population

Job title and employment duration

SMR, Mantel-Haenszel RR

Age, race

Paci et al., 1989

Mortality experience (1939–1984) of shoe workers (at least 1 day) in Florence, Italy

2013 total

1008 men

1005 women

Italy general population

Plant production records and work histories

SMR

Age, sex, calendar year

Walker et al., 1993

Mortality experience (1940–1982) of shoe-manufacturing workers (at least 1 month in 1940–1979) in Ohio

7814 total

2529 men

5285 women

US general population

Employment at one of two plants

SMR

Age, sex, race, time period

 

Other exposures: MEK, acetone, naphtha, isopropyl alcohol, methanol, ethylene glycol monoethyl ether, xylene

 

Greenland et al., 1994

White, male cases of cancer (multiple sites; died in 1969–1984) and controls in a cohort of transformer-assembly workers in Massachusetts

1821 cases

1202 controls

Internal comparison

Job titles rated for exposure by industrial hygienist

Logistic OR (nested case-control)

 

Other exposures: trichloroethylene, solvents

Age, death year, covariates that altered an estimate >20%

Lagorio et al., 1994

Mortality experience (1981–1992) of self-employed gas-station attendants (in 1980) in Italy

2665 total

2308 men

357 women

Latium region, Italy general population

Environmental survey and duration of employment

SMR

Age, sex

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description

Study Group (N)

Comparison Group (N)

Exposure Assessment and Other Relevant Exposures

Analysis and Adjustment for Potential Confounders

Heineman et al., 1995

Brain tumor incidence in women (1980–1984) in Shanghai, China

276

Shanghai general population

Job title

SIR

Age

Fu et al., 1996

Mortality experience (1939–1991) of shoe-manufacturing workers in England (1939) and Italy (1950–1984)

6223 total

5220 men

1003 women

England and Italy general populations

Job title

SMR

Age, sex, time period

Schnatter et al., 1996a,b

Cases of lymphohematopoietic cancers (died in 1964–1983) and controls in a cohort of Canadian petroleum-distribution workers

29 cases, matched 1:4

Internal comparison

Industrial hygienist review based on work histories, site characterizations, surveys

Mantel-Haenszel OR (nested case-control)

Smoking, family cancer history, x-ray history

Ireland et al., 1997

Mortality experience (1940–1991) of male US chemical-plant workers (at least 1 day in 1940–1977) in Monsanto company plant in Sauget, IL

4172

Illinois general population

Industrial hygienist exposure estimates based on work records

SMR

Lynge et al., 1997

Incidence experience (1970–1991) in service-station workers (1970) in Denmark and Scandinavia

18,969 total

16,524 men

2445 women

Nation general populations

Job title

SMR

Age, sex

Rushton and Romaniuk, 1997

Cases of leukemia and controls in a cohort of petroleum-distribution workers (1975–1992) in UK

91 cases, matched 1:4

Internal comparison

Measurements factored in occupational hygiene estimates, work histories, job descriptions, fuel compositions

OR (nested case-control), logistic regression

Age, smoking, date of hire, employment duration, socioeconomic status

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description

Study Group (N)

Comparison Group (N)

Exposure Assessment and Other Relevant Exposures

Analysis and Adjustment for Potential Confounders

Trichloroethylene

Aircraft and Aerospace Workers

Garabrant et al., 1988

Mortality experience (1958–1982) of aircraft-manufacturing workers (at least 1 day) at an aircraft-manufacturing facility in San Diego County, California (with at least 4 years of cumulative company employment)

14,067 total

11,898 men

2169 women

US general population

Employment determined through company work records and interviews

SMR

Age, sex, race, calendar year, duration of employment, year of death

Spirtas et al., 1991

Mortality experience (1952–1982) of aircraft-maintenance workers (at least 1 year in 1952–1956) at Hill Air Force Base in Utah

14,457 total

10,730 men

3727 women

Utah white population

Industrial hygienist assessment from interviews, surveys, hygiene files, position descriptions

SMR, trend analysis

Age, sex, calendar period

 

Other exposures: Stoddard solvent, isopropyl alcohol, trichloroethane, acetone, toluene, MEK, methylene chloride

 

Blair et al., 1998

Incidence and mortality experience (1952–1990) of aircraft-maintenance workers (at least 1 year in 1952–1956) at Hill Air Force Base in Utah

14,457 total

10,730 men

3727 women

Utah white population

Industrial hygienist assessment from interviews, surveys, hygiene files, position descriptions

SMR, RR (Poisson)

Age, sex, calendar period

 

Other exposures: Stoddard solvent, isopropyl alcohol, trichloroethane, acetone, toluene, MEK, methylene chloride

 

Morgan et al., 1998

Mortality experience (1950–1993) of aerospace workers (at least 6 months) Hughes Aircraft plant in Arizona

20,508 total

(4733 exposed)

13,742 men

6766 women

US general population

Exposure matrixes generated by employees and industrial hygienists

SMR, Cox proportional hazards model

Age, sex

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description

Study Group (N)

Comparison Group (N)

Exposure Assessment and Other Relevant Exposures

Analysis and Adjustment for Potential Confounders

Boice et al., 1999

Mortality experience (1960–1996) of aircraft-manufacturing workers (at least 1 year) Lockheed Martin facility in California

77,965 total

62,477 men

15,488 women

General California population of white workers

Abstracted from walkthrough surveys, hygiene files, job descriptions

SMR, RR (Poisson)

Age, sex, race, dates of first and last employment

 

Other exposures: tetrachloroethylene, solvents

Other Cohort Studies

Axelson et al., 1978

Mortality experience (1955–1975) of Swedish men occupationally exposed during the 1950s and 1960s

518

Sweden general population

Biologic monitoring for U-TCA

RR

Age

Axelson et al., 1994

Mortality experience (1955–1986) of Swedish workers occupationally exposed during the 1950s and 1960s

1670 total

1421 men

249 women

Sweden general population

Biologic monitoring for U-TCA

SMR, SIR (Poisson)

Age, sex, time period

Anttila et al., 1995

Incidence experience (1967–1992) of workers biologically monitored for occupational exposure to halogenated solvents (1965–1983) at the Finnish Institute of Occupational Health

3974 total

2050 men

1924 women

Finland general population

Biologic monitoring for U-TCA, and blood metabolites of tetrachloroethylene and trichloroethane

SIR

Age, sex, time period

 

Other exposures: trichloroethane, tetrachloroethylene

 

Ritz, 1999

Mortality experience (1951–1989) of male uranium-processing plant workers (at least 3 years, with first hire in 1951–1972) in Ohio

3814

(1) External comparison with US general population

(2) Internal comparison among workers monitored for exposure

Exposure matrixes generated by employees and industrial hygienists

SMR, RR (conditional logistic regression)

Age, calendar year, time since first hired, pay type, radiation dose

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description

Study Group (N)

Comparison Group (N)

Exposure Assessment and Other Relevant Exposures

Analysis and Adjustment for Potential Confounders

Hansen et al., 2001

Incidence experience (1968–1996) in Danish workers (1947–1989) occupationally exposed

803 total

658 men

145 women

Denmark general population

Biologic monitoring for U-TCA

SIR

Age, sex, calendar year, period of first employment, employment duration

Tetrachloroethylene

Dry-cleaning Cohorts

Brown and Kaplan, 1987

Mortality experience (1960–1982) of dry cleaners (at least 1 year, before 1960) in four US labor unions

1690

US general population

Employment in dry-cleaning shops using tetrachloroethylene or other solvents

SMR

Age, time period

Ruder et al., 1994

Mortality experience (1960–1990) of dry cleaners (at least 1 year, before 1960) in four US labor unions

1701 total

592 men

1109 women

SMRs calculated with modified life-table analysis system of NIOSH

Employment in dry-cleaning shops using tetrachloroethylene or other solvents

SMR

Age, sex, time period

Ruder et al., 2001

Mortality experience (1960–1996) of dry cleaners (at least 1 year, before 1960) in four US labor unions

1701 total

592 men

1109 women

SMRs calculated with modified life-table analysis system of NIOSH

Employment in dry-cleaning shops using tetrachloroethylene or other solvents

SMR

Age, sex, time period

Blair et al., 1990

Mortality experience (1948–1978) of members of a dry-cleaning union in St. Louis

5365 total

1319 men

4046 women

US general population

Exposure index created from job title and length of union membership

SMR, trend analysis

Age, sex, calendar year, race

Other Cohort Studies

Lynge and Thygesen, 1990

Incidence experience (1970–1980) of Danish laundry and dry-cleaning workers (in 1970)

10,600 total

2033 men

8567 women

Denmark general population

Dry-cleaning, job title

SIR

Age

 

Other exposures: trichloroethylene, 1,1,2-trichloro-1,2,2-trifluoroethane

 

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description

Study Group (N)

Comparison Group (N)

Exposure Assessment and Other Relevant Exposures

Analysis and Adjustment for Potential Confounders

Anttila et al., 1995

Incidence experience (1967–1992) of workers biologically monitored for occupational exposure to halogenated solvents (1965–1983) at the Finnish Institute of Occupational Health

3974 total

2050 men

1924 women

Finland general population

Biologic monitoring

SIR

Age, sex, time period

 

Other exposures: trichloroethane, trichloroethylene

 

Boice et al., 1999

Mortality experience (1960–1996) of aircraft-manufacturing workers (at least 1 year) at Lockheed Martin facility in California

77,965 total

62,477 men

15,488 women

General California population of white workers

Abstracted from walkthrough surveys, hygiene files, job descriptions

SMR, RR (Poisson)

Age, sex, race, dates of first and last employment

 

Other exposures: trichloroethylene, solvents

Methylene Chloride

Kodak Park Cohort

Friedlander et al., 1978

PMR of former or current exposed workers (1956–1976) at Kodak Park

334

Deaths of former or current unexposed workers (1956–1976) at Kodak Park

Employment in methylene chloride area

PMR, SMR

Age, sex

Mortality experience (1964–1976) of hourly-wage male workers (in 1964) at Kodak Park

751

Hourly-wage male workers at Kodak Park and in New York state (excluding New York City) men

 

Hearne and Friedlander, 1981

Mortality experience (1964–1980) of hourly-wage male workers (in 1964) at Kodak Park

750

Hourly-wage male workers at Kodak Park and in New York state (excluding New York City) men

Employment in methylene chloride area

SMR

Age, sex

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description

Study Group (N)

Comparison Group (N)

Exposure Assessment and Other Relevant Exposures

Analysis and Adjustment for Potential Confounders

Hearne et al., 1987

Mortality experience (1964–1984) of hourly-wage male workers (in 1964) at Kodak Park

1013

(1) New York state (excluding New York City) men (1945–1990)

(2) Over 40,000 Rochester-based Kodak workers

Employment in roll-coating division, cumulative exposure assigned by industrial hygienist from historical records

SMR

Age, sex, time period

Hearne et al., 1990

Mortality experience (1964–1988) of hourly-wage male workers (in 1964) at Kodak Park

1013

(1) New York state (excluding New York City) men (1945–1990)

(2) Over 40,000 Rochester-based Kodak workers

Employment in roll-coating division, cumulative exposure assigned by industrial hygienist from historical records

SMR, trend analysis

Age, sex, time period

Hearne and Pifer, 1999

Mortality experience (1946–1994) of two overlapping cohorts of exposed male workers (at least 1 year in 1946–1970; any employment in 1964–1970) at Kodak Park

(1) 1311

(2) 1013

New York state (excluding New York City) men (1945–1990)

Employment in methylene chloride area, cumulative exposure assigned by industrial hygienist from historical records

SMR, trend analysis

Age, sex, time period

Cellulose-Fiber Production Cohort

Ott et al., 1983

Mortality experience (1954–1977) of cellulose-fiber production plant workers (at least 3 months) in Rock Hill, SC

1271 total

551 men

720 women

York County, SC, general population

Employment in plant, comprehensive industrial hygienist survey

SMR, conditional risk, Cox regression

Age, sex, race, year of first exposure

 

Other exposures: acetone, methanol

Lanes et al., 1990

Mortality experience (1954–1986) of cellulose-fiber production plant workers (at least 3 months in 1954–1977) in Rock Hill, SC

1271 total

551 men

720 women

York County, SC, general population

Employment in plant, comprehensive industrial hygienist survey

SMR

Age, sex, race

 

Other exposures: acetone, methanol

 

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description

Study Group (N)

Comparison Group (N)

Exposure Assessment and Other Relevant Exposures

Analysis and Adjustment for Potential Confounders

Lanes et al., 1993

Mortality experience (1954–1990) of cellulose-fiber production plant workers (at least 3 months in 1954–1977) in Rock Hill, SC

1271 total

551 men

720 women

York County, SC, general population

Employment in plant and comprehensive industrial hygienist survey

SMR

Age, sex, race

 

Other exposures: acetone, methanol

 

Other Cohort Studies

Gibbs et al., 1996

Mortality experience (1970–1989) of cellulosefiber production workers (at least 3 months) in Cumberland, MD

3211 total

2187 men

1024 women

Allegheny County, MD, general population

Workplace monitoring data, job title

SMR

Age, sex, race, time period

Tomenson et al., 1997

Mortality experience (1946–1994) of male cellulose triacetate film-base workers (any employment in 1946–1988) in Brantham, UK

1785

England and Wales mortality rates

Workplace monitoring

SMR, trend analysis

Age, time period

Toluene and Xylene

Swedish Paint Industry Cohort

Lundberg, 1986

Incidence and mortality experience (1955–1981) of male, Swedish paint-industry workers (at least 5 years in 1955–1975) with long-term exposure to organic solvents

416

Sweden general population

Industry employment, historical air exposure measurements

SMR

Sex, time period

Lundberg and MilatouSmith, 1998

Incidence and mortality experience (1955–1994) of male, Swedish paint-industry workers (at least 5 years in 1955–1975) with long-term exposure to organic solvents

411

Sweden general population

Industry employment, historical air exposure measurements

SMR, SIR

Age, sex, time period

Other Cohort Studies

Anttila et al., 1998

Incidence experience (1973–1992) in workers biologically monitored for occupational exposure to aromatic hydrocarbons (1973–1983) at the Finnish Institute of Occupational Health

5301 total

3922 men

1379 women

Finland general population

Biologic monitoring

SIR

Age, sex, time period

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description

Study Group (N)

Comparison Group (N)

Exposure Assessment and Other Relevant Exposures

Analysis and Adjustment for Potential Confounders

Svensson et al., 1990

Incidence and mortality experience (1925–1985) of Swedish male rotogravure workers (at least 3 months)

1020

Population-specific rates for geographic area around factory

Exposures evaluated through plant visits, biologic monitoring, workplace measurements, historical documents, interviews

SMR, SIR

Age, sex, calendar year, location

 

Other exposures: benzene, solvents

 

Solvents

UK Rubber Worker Cohort

Parkes et al., 1982

Mortality experience (1946–1975) of UK male rubber workers (at least 1 year in 1946–1960)

33,815

UK general population

Industry employment

SMR

Age, sex

Sorahan et al., 1986

Mortality experience (1946–1980) of UK male rubber workers (at least 1 year in 1946–1960)

36,445

UK general population

Industry employment

SMR, regression models and life tables

Age, sex, age at hire, entry cohort, location, work sector

Sorahan and Cathcart, 1989

Mortality experience (1946–1985) of UK male rubber workers (at least 1 year in 1946–1960)

36,691

UK general population

Industry employment

SMR, regression models and life tables

Age at hire, entry cohort, location, work sector, duration of employment

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description

Study Group (N)

Comparison Group (N)

Exposure Assessment and Other Relevant Exposures

Analysis and Adjustment for Potential Confounders

Other Cohort Studies

Costantini et al., 1989

Mortality experience (1950–1983) of male workers at tanneries (at least 6 months) in Tuscany, Italy

2926

Italy general population

6 months of employment

SMR

Age, time period

Guberan et al., 1989

Incidence and mortality experience (1970–1984) of painters and electricians (in 1970) in the Canton of Geneva

1916 painters

1948 electricians

Switzerland regional population

Job title

SMR, SIR

Age

Acquavella et al., 1993

Mortality experience (1950–1987) of workers hired at a metal components manufacturing facility (at least 6 months in 1950–1967) in Iowa

3630 total

2664 men

966 women

Iowa general population

Occupational titles, departments

SMR, RR

Age, sex, time period

Berlin et al., 1995

Incidence and mortality experience (1967–1987) of Swedish workers occupationally exposed to solvents

5791 total

5283 men

508 women

Sweden general population

Patients with solvent-related disorders

SMR, SIR

Age, sex

Lynge et al., 1995

Incidence experience (1970–1987) in Danish printing workers (in 1970)

19,127 total

15,534 men

3593 women

Economically active people in Denmark

Job title

SIR

Age, alcohol and tobacco use

Steenland and Palu, 1999

Mortality experience (through 1994) of members of US painters unions (at least 1 year of membership; born before 1940) in four states

42,170

US general population, nonpainter cohort

Union membership

SMR, SRR

Age, time period

Other Specific Solvents

Isopropyl Alcohol and Methyl Ethyl Ketone

Alderson and Rattan, 1980

Mortality experience (1935–1975) of workers in Shell MEK dewaxing or isopropyl alcohol plants (at least 1 year) in Britain

262 IAP

446 MEK

US general population

Employment in one of two plants

SMR

Age, time period

Phenol

Dosemeci et al., 1991

Mortality experience (1966–1979) of white male workers (employed before 1966) employed at five facilities producing or using phenol and formaldehyde

14,861

US general population

Employment at one of five facilities

SMR

Age, time period

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description

Study Group (N)

Comparison Group (N)

Exposure Assessment and Other Relevant Exposures

Analysis and Adjustment for Potential Confounders

Ethanol and Isopropyl Alcohol

Teta et al., 1992

Mortality experience (1940–1983) of male isopropanol and ethanol production workers (1940–1978) at two facilities in South Charleston, WV, and Texas City, TX

1031

(1) White subjects and South Charleston cohort compared with general US white male population

(2) Nonwhite subjects from Texas compared with US nonwhite male population

Employment in one of two plants

SMR

Age, time period, duration of assignment, time since first assignment, year of first assignment, type of assignment

NOTE: SMR=standardized mortality ratio; RR=relative risk; SIR=standardized incidence ratio; OR=odds ratio; PMR=proportional mortality ratio; SRR=standardized relative risk; NIOSH=National Institute for Occupational Safety and Health; IAP=isopropyl alcohol; MEK=methyl ethyl ketone; U-TCA=urinary metabolite of trichloroethylene.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Although the association between benzene and cancer was assessed in a number of cohort studies, two studies provide the most comprehensive data. The first was designed to investigate mortality in three Ohio rubber-manufacturing plants, referred to as the “Pliofilm” cohort (Crump, 1994, 1996; Infante et al., 1977; Paxton, 1996; Paxton et al., 1994a,b; Rinsky et al., 1987; Wong, 1995). Benzene was used in the production of rubber hydrochloride, a natural rubber cast film used primarily for wrapping foods and marketed under the trade name Pliofilm. The plants were chosen because of the relatively high exposure to benzene and the lack of other toxic chemicals in use. In 1975, after a report of several leukemia cases, the National Institute for Occupational Safety and Health conducted a retrospective study of 748 Pliofilm workers who were exposed to benzene (Infante et al., 1977). Rinsky and colleagues (1981, 1987) expanded on the work of Infante and colleagues by increasing the size of the cohort to 1006 workers, extending the years of observation, and collecting additional exposure data from the plants’ processes, company records, and air-sampling data (Rinsky et al., 1981), thereby providing estimates of exposure for each job in the various areas of the plant (Rinsky et al., 1987). The retrospective exposure assessment was further modified by Crump and Allen (1984) and Paustenbach and colleagues (1992). Crump and Allen (1984) developed an exposure matrix based on the concept that the benzene levels in the workplace may have improved over time, whereas the exposure matrix developed by Paustenbach and colleagues (1992) incorporated more detailed information from monitoring devices, the changing length of the workweek over the years, the impact of World War II on production, engineering controls, and other available and experimental data to assess exposure to benzene. Considerable controversy surrounds the assessments of exposure (Wong, 1995). Different authors have used the different exposure assessments in their analyses: the Rinsky exposure assessment (Paxton et al., 1994a,b; Rinsky et al., 1981, 1987), the Crump assessment (Paxton et al., 1994a,b), and the Paustenbach assessment (Crump, 1994; Paxton et al., 1994a,b). The differences in the exposure assessments lead to differences in the estimates of relative risks for various sites of cancer according to exposure to benzene. The differences in relative risk are important in setting regulatory standards but were not sufficiently different to affect the committee’s determination of the magnitude of association.

The second key study of occupational exposure to benzene was conducted in China. After conducting a nationwide benzene-monitoring survey in China during 1979–1981, Yin and colleagues at the Chinese Academy of Preventive Medicine identified a cohort of about 30,000 workers who were occupationally exposed to benzene or mixtures containing benzene (Yin et al., 1987). Subjects were selected from painting, shoe-making, rubber synthesis, leather, and adhesive and organic-chemical synthesis factories. A sample of 28,257 workers employed in machine production, textile, and cloth factories was taken to represent an unexposed comparison population. Later studies expanded the original cohort to 74,828 benzene-exposed and 35,805 nonexposed workers and included a detailed assessment of exposure to benzene. Those studies were conducted in collaboration with the US National Cancer Institute (Hayes et al., 1996, 1997; Li et al., 1994; Yin et al., 1987, 1989, 1994, 1996a,b). Like the Pliofilm study, studies of the cohort yielded valuable information regarding the risk of developing or dying from cancer in relation to exposure to benzene.

Other important cohort studies of benzene-exposed workers include those of American chemical workers (Ireland et al., 1997; Wong, 1987a,b), female workers in China

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

(Heineman et al., 1995), other rubber-plant workers (McMichael et al., 1976; Wilcosky et al., 1984), shoe-manufacturing workers (Fu et al., 1996; Paci et al., 1989; Pippard and Acheson, 1985; Walker et al., 1993), transformer-assembly workers (Greenland et al., 1994), filling and service-station attendants (Lagorio et al., 1994, Lynge et al., 1997), and petroleum distributors (Rushton and Romaniuk, 1997; Schnatter et al., 1996a,b). These studies differed substantially from the two preceding studies in that exposure was much lower. For example, levels in the Pliofilm cohort ranged from 7.2 to 24.9 ppm (Rinsky et al., 1981, 1987), whereas levels in the cohort of petroleum workers ranged from 0.01 to 6.2 ppm (Ireland et al., 1997; McMichael et al., 1976; Rushton and Romaniuk, 1997; Schnatter et al., 1996a,b).

Studies of Workers Exposed to Trichloroethylene

The most important use of trichloroethylene has been in the removal of greases, tars, and oils from metal parts. It has also been used by the textile industry to scour cotton, wool, and other fabrics and as a solvent in waterless dying and finishing operations (ATSDR, 1997b). The regulatory limit set by OSHA is a 8-hr TWA of 100 ppm (NIOSH, 1997).

Mortality in relation to exposure to trichloroethylene has been examined in four large cohort studies of aircraft and aerospace manufacturing and maintenance workers (Blair et al., 1998; Boice et al., 1999; Garabrant et al., 1988; Morgan et al., 1998). In general, industrial hygienists reviewed information obtained from walkthrough surveys, interviews of long-term employees, and historical information on job titles and tasks, operations, and worksites to classify workers by duration and intensity of exposure. The first study was conducted to evaluate mortality rates among 14,457 aircraft maintenance workers at Hill Air Force Base, Utah, in response to concerns expressed by workers in the middle-1970s about potential health effects of chemical exposure (Spirtas et al., 1991). Trichloroethylene was used as a vapor degreaser until 1978, when it was replaced with 1,1,1-trichloroethane. It was also used to clean small electric parts at work benches until 1968 (Blair et al., 1998). Other solvents used at the base were primarily other chlorinated hydrocarbons, aromatic hydrocarbons, and carbon tetrachloride. Blair and co-workers (1998) extended the followup of the cohort assembled by Spirtas and colleagues (1991) by 8 years. The estimates of exposure developed for the initial cohort study (Stewart et al., 1991) were also used in the extended followup study.

Two additional large cohort studies evaluated mortality in aircraft manufacturing facilities where trichloroethylene was commonly used as a degreaser. Boice and colleagues (1999) investigated 77,965 workers at Lockheed Martin’s Burbank California factories, and Morgan and co-workers (1998) reported on 20,508 employees at a Hughes Aircraft manufacturing facility in Arizona.

Other important but smaller cohort studies of workers exposed to trichloroethylene are those of Swedish trichloroethylene production workers (Axelson et al., 1978, 1994), US uranium processing-plant workers (Ritz, 1999), and other workers occupationally exposed to trichloroethylene in Finland (Anttila et al., 1995) and Denmark (Hansen et al., 2001). Exposure-response analyses in the Scandinavian studies were based on biologic monitoring of urinary-trichloroacetic acid (U-TCA, a metabolite of trichloroethylene) measured in urine samples from workers (Anttila et al., 1995; Axelson et al., 1978, 1994; Hansen et al., 2001). Hansen and colleagues (2001) also used data on levels of trichloroethylene in the breathing

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

zone of workers. The US study used semiquantitative exposure estimates based on expert review (Ritz, 1999).

Studies of Workers Exposed to Tetrachloroethylene and Dry-cleaning Solvents

Tetrachloroethylene has been used for metal cleaning and vapor degreasing and for dry-cleaning and textile processing. The PEL set by OSHA is a 8-hr TWA of 100 ppm (NIOSH, 1997).

Occupational exposure to tetrachloroethylene has been studied primarily in dry-cleaning workers because of its widespread use. Dry-cleaning workers are extensively exposed to organic solvents, which are integral to the dry-cleaning process. The evolution of the dry-cleaning process has seen the use of a variety of solvents. Most of the early dry-cleaning solvents were petroleum-based and included naphtha and Stoddard solvent. The petroleum-based solvents were replaced in the 1930s largely with carbon tetrachloride, a less expensive alternative (IARC, 1995). Information about the toxicity and corrosiveness of carbon tetrachloride led to its replacement in the 1950s with chlorinated hydrocarbons. Today, tetrachloroethylene is the most commonly used dry-cleaning solvent in the United States. Other solvents and chemicals used in dry-cleaning include 1,1,2-trichloro-1,2,2-trifluoroethane, and 1,1,1-trichloroethane (IARC, 1995).

An important study examined a cohort of 1708 US dry-cleaning workers drawn from four labor unions, first reported by Brown and Kaplan (1987), and updated by Ruder and colleagues (1994, 2001). The original study investigated mortality through 1982, the first update extended the followup through 1990 (Ruder et al., 1994), and the most recent study updated mortality through 1996 (Ruder et al., 2001). In the updates, two subcohorts were evaluated on the basis of employment in shops where tetrachloroethylene was the cleaning solvent (625 workers) or in shops where tetrachloroethylene use could not be confirmed or another solvent was used as the cleaning solvent (1083 workers).

Another study, of 5365 members of a dry-cleaning union in Missouri, assessed mortality in relation to estimated levels of exposure to dry-cleaning solvents (Blair et al., 1990). Exposure indexes were based on job title and type of establishment. Information on the type of solvent used was not available so workers who specifically used tetrachloroethylene could not be identified.

Other studies of exposure to tetrachloroethylene and cancer include a US study of aircraft manufacturing workers (Boice et al., 1999) and a Finnish study of workers occupationally exposed to three halogenated hydrocarbons, including tetrachloroethylene (Anttila et al., 1995). The exposure assessment of the US study was based on expert review of walkthrough surveys and historical documents and other approaches (Boice et al., 1999), whereas the Finnish study estimated level of exposure from biologic monitoring (Anttila et al., 1995). Another study examined Danish laundry and dry-cleaning workers whose chemical exposure was inferred from their occupations, as specified by census industry codes (Lynge and Thygesen, 1990).

Studies of Workers Exposed to Methylene Chloride

Methylene chloride (dichloromethane) has been used in degreasing, in paint stripping, as an aerosol propellant, and in the manufacture of textiles, plastics, and photographic film. A large proportion of workers exposed to methylene chloride are

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

involved in metal cleaning, industrial paint stripping, and using ink solvents (ATSDR, 2000). The regulatory limits have decreased as information on toxicity has accumulated (ATSDR, 2000). The current PEL set by OSHA is a 8-hr TWA of 25 ppm (NIOSH, 1997).

The key occupational cohort study of exposure to methylene chloride was an incidence and mortality study of workers in an Eastman Kodak plant (Friedlander et al., 1978; Hearne and Friedlander, 1981; Hearne and Pifer, 1999; Hearne et al., 1987, 1990). The workers, ranging in number from 750–1311, were exposed chronically to methylene chloride in the manufacturing of cellulose triacetate, a photographic film base, as confirmed by study personnel who used air sampling and gas chromatography. In the most recent update, Hearne and Pifer (1999) followed the mortality experience through 1994 for two groups of workers: 1311 workers who first worked in film-support manufacturing and related operations in 1946–1970 and the Roll Coating Cohort (1964–1970) of 1013 employees that was previously studied (Hearne et al., 1987, 1990). Exposure to methylene chloride was estimated by combining air-monitoring data with information on work histories.

Other important studies include a study of workers employed at a plant that produced cellulose triacetate film base in the UK (Tomenson et al., 1997), and studies of workers in the production of cellulose fiber at a Hoechst manufacturing plant in South Carolina (Lanes et al., 1990, 1993; Ott et al., 1983) and a Hoechst plant in Maryland (Gibbs et al., 1996). Estimates of exposure in the UK cohort were derived from area-monitoring results, work histories, and historical information on production processes (Tomenson et al., 1997). Some exposure-monitoring data were obtained on the South Carolina cohort in the 1970s (Ott et al., 1983), but exposure estimates were unavailable for most of the study period. Exposure measurements were not used in the mortality analysis of the South Carolina cohort (Lanes et al., 1990, 1993), but Gibbs and co-workers (1996) used the available exposure data to determine high and low exposure ranges.

Studies of Workers Exposed to Other Specific Solvents

Three studies of solvent-production plants were used to evaluate mortality in relation to exposure to specific solvents: isopropanol (Alderson and Rattan, 1980; Teta et al., 1992), methyl ethyl ketone (Alderson and Rattan, 1980), phenol (Dosemeci et al., 1991), and toluene, xylene, and styrene (Anttila et al., 1998). The main uses of isopropanol are as a chemical intermediate and in applications in medicine and industry (Logsdon and Loke, 1996). The PEL is a 8-hr TWA of 400 ppm (NIOSH, 1997). Methyl ethyl ketone is used primarily as a solvent in industry. The regulatory limit set by OSHA is 200 ppm (NIOSH, 1997). Phenol is commonly used in the production of epoxy resins and polycarbonates, phenolic resins and molding compounds, caprolactam, aniline alkylphenols, and xylenols; as a fungicide or disinfectant; and in a variety of medications (ATSDR, 1998; Wallace, 1996). The occupational exposure limit set by OSHA is a 8-hr TWA of 5 ppm (NIOSH, 1997). Toluene and xylene are used in the manufacture of a variety of chemicals and as solvents for paints, lacquers, gums, printing inks, and resins. Styrene is used in the production of polystyrene plastics and resins and as an intermediate in the production of such copolymers as styrene-acrylonitrile, acrylonitrile-butadiene-styrene, and styrene-butadiene rubber. The occupational regulatory limit for toluene is a 8-hr TWA of 200 ppm (NIOSH, 1997), and the PEL for xylene and styrene is a 8-hr TWA of 100 ppm (NIOSH, 1997).

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

The key studies for evaluating risks posed by those solvents include the following. Alderson and Rattan (1980) evaluated mortality in 262 male workers employed in isopropanol plant and 446 male employees of two methyl ethyl ketone dewaxing plants, using the type of plant as an indicator of exposure. Teta and colleagues (1992) conducted a cohort mortality study of 1031 workers employed at two facilities that produced ethanol and isopropanol. Employment in an isopropanol strong-acid production unit was used as an exposure surrogate for isopropanol. Dosemeci and co-workers (1991) conducted a mortality followup of 14,861 workers employed in five plants that manufactured or used phenol and formaldehyde. Estimates of exposure to phenol were derived from expert review of information obtained from walkthrough survey reports, historical monitoring results, and other workplace information. Anttila and colleagues (1998) investigated 3922 male and 1379 female Finnish workers occupationally exposed to toluene, xylene, and styrene; level of exposure was determined from biologic monitoring (Anttila et al., 1995).

Studies of Workers Exposed to Unspecified Mixtures of Organic Solvents

Solvents are used in numerous occupations, so the committee examined cancer mortality and incidence in workers in a number of occupations that may have involved exposure to organic solvents. It is important to note that workers in the occupations in question have potential exposure to numerous chemicals in addition to solvents.

Painters have the potential for frequent and high level exposure to many types of organic solvents. Organic solvents, such as, toluene, xylene, glycols, and methylene chloride, have been used over the years in the composition of paint, paint thinners, cleaners, and strippers. Painters are exposed to numerous other chemical and environmental agents, including pigments, dusts, resins, and silicates. In most instances, it was not possible to identify which solvents were used, and the committee referred to them as unspecified mixtures of organic solvents. Lundberg and Milatou-Smith (1998) evaluated cancer mortality and incidence in a cohort of 411 male workers who had been employed for more than 5 years in 1955–1975 in the Swedish paint-manufacturing industry (followup of Lundberg, 1986). Guberan and colleagues (1989) studied cancer mortality and incidence in 1916 painters in Geneva, Switzerland, who were identified from the 1970 census. The largest cohort study of painters was conducted by Steenland and Palu (1999), who evaluated mortality patterns in a cohort of 42,170 painters who were members of the Painters Union for 1 year or more before 1979.

Other cohorts exposed to unspecified chemical mixtures involve printers (Lynge et al., 1995; Svensson et al., 1990); workers in solvent-production plants (Berlin et al., 1995); metal workers (Acquavella et al., 1993); rubber workers (Parkes et al., 1982; Sorahan and Cathcart, 1989; Sorahan et al., 1986); workers in tanneries (Costantini et al., 1989); and shoemakers (Fu et al., 1996; Paci et al., 1989; Pippard and Acheson, 1985; Walker et al., 1993). In the past, some shoemaking cohorts would have had considerable exposure to benzene (e.g., Fu et al., 1996; Paci et al., 1989). However, because the composition of glues has changed, benzene being replaced with other solvents, and because there were no precise estimates of exposure, the committee classified those cohorts as being exposed to unspecified mixtures of organic solvents.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

ORAL, NASAL, AND LARYNGEAL CANCER

Description of Case-Control Studies

Two case-control studies reviewed by the committee that included oral, nasal, or laryngeal cancers are described in Table 6.2. One population-based case-control study examined the risk of nasal and nasopharyngeal cancer associated with occupational exposure to organic solvents (Hardell et al., 1982), and a second examined the risk of oral cavity and laryngeal cancers associated with work in the dry-cleaning industry (Vaughan et al., 1997). Both included interviews with study subjects concerning occupational history. In the former study, exposure to organic solvents was self-reported; in the latter, levels of exposure to tetrachloroethylene in cleaning jobs were assigned by an industrial hygienist.

Epidemiologic Studies of Exposure to Organic Solvents and Oral Cancer

Three cohort studies and one case-control study failed to provide strong evidence of an association between tetrachloroethylene and dry-cleaning solvents and oral cancer. Ruder and colleagues (2001) found a strong, increased risk of cancer of the tongue (standardized mortality ratio [SMR]=9.03, 95% confidence interval [CI]=1.86–26.39) in the subcohort exposed only to tetrachloroethylene. In another cohort of dry cleaners, Blair and colleagues (1990) found no increased risk of cancers of the buccal cavity and pharynx (SMR=1.0, 95% CI=0.3–2.2).

The case-control study undertaken by Vaughan and colleagues (1997) found little evidence of an increased risk of oral cancer in dry-cleaning workers (odds ratio [OR]possible exposure=1.2, 95% CI=0.3–4.6; ORprobable exposure=1.5, 95% CI=0.2–9.5). The relative risks increased with increasing cumulative exposure but not with duration of employment, although there was considerable statistical uncertainty in the trend.

In only one study was the risk of cancers of the mouth and throat evaluated among workers exposed to phenol (Dosemeci et al., 1991); no increased risk was found (SMR=0.8, 95% CI=0.4–1.5).

The risk of oral cancer in industries exposed to solvents was estimated in cohorts of workers in cellulose-fiber production (Lanes et al., 1990), shoe manufacture (Walker et al., 1993), methyl ethyl ketone dewaxing (Alderson and Rattan, 1980), and ethanol and isopropanol production (Teta et al., 1992). Although specific solvents were used in most of those occupations, many other solvents were also used, and specific solvents were not evaluated in any of the studies. There was no evidence of positive associations in any of these studies.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.2 Description of Case-Control Studies of Oral, Nasal, and Laryngeal Cancer and Exposure to Organic Solvents

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Hardell et al., 1982

Male cases, age 28–85 years, reported to the Swedish Cancer Registry in 1970–1979 from the three northernmost counties of Sweden; controls from authors’ previous studies on soft tissue sarcoma and malignant lymphoma were used; controls identified from the Swedish National Population Registry and the National Registry for Causes of Death

44 nasal

27 nasopharyngeal

541

Organic solvents

Occupational histories (job titles) and exposures (self-reported) obtained through questionnaire and supplemented by telephone interview (direct or proxy)

Exposure frequency

None

Vaughan et al., 1997

Cases, age 20–74 years when diagnosed, identified from a cancer surveillance system covering 13 counties in western Washington; cases lived in one of the three largest counties and were diagnosed in 1983–1990; population-based controls, frequency-matched by age and sex, identified through RDD

491 oral cavity

235 laryngeal

724

Tetrachloroethylene

Dry-cleaning work

In-person interviews to assess employment and duration in dry-cleaning occupations; probability of exposure from decade of employment; cumulative exposure measurements calculated by duration and occupation-specific time-weighted averages

Conditional logistic regression

Age, sex, education, study period, alcohol consumption, cigarette smoking, race

Response rates: 85.2% of oral cavity cancer cases, 80.8% of laryngeal cancer cases, 80.3% of controls

NOTE: RD=random-digit dialing.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Summary and Conclusion

There was little consistent evidence of an association between oral cancer and exposure to tetrachloroethylene and dry-cleaning solvents. Although there were several positive studies, most were based on small numbers of exposed cases and did not have sufficient statistical power. For exposure to phenol, only one study was identified and a risk of oral cancer was not found. No risk was also found among the occupational studies on solvent mixtures. Table 6.3 identifies the studies reviewed on oral cancer. Unless indicated in the table, the study populations include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to solvents under review and oral cancer.

TABLE 6.3 Selected Epidemiologic Studies—Oral Cancer and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Specific and Unspecified Mixtures of Organic Solvents

Cohort Studies—Mortality

Ruder et al., 2001

Dry-cleaning union workers exposed to tetrachloroethylene

 

 

Cancer of the tongue

3

9.03 (1.86–26.39)

Walker et al., 1993

Ohio shoe-manufacturing employees

 

 

Males and females

5

0.67 (0.22–1.59)

Teta et al., 1992

Male workers at ethanol/isopropanol production plants

 

 

South Charleston

2

1.3 (0.2–4.8)a

 

Texas City

1

1.4 (0.0–8.4)a

Dosemeci et al., 1991

Male industrial workers exposed to phenol

11

0.8 (0.4–1.5)

Blair et al., 1990

St. Louis, MO, dry-cleaning workers

5

1.0 (0.3–2.2)

Lanes et al., 1990

Cellulose-fiber production workers

2

2.31 (0.39–7.60)

Alderson and Rattan, 1980

Male British workers at two methyl ethyl ketone dewaxing plants

2

15.38 (1.86–55.54)a

Case-Control Study

Vaughan et al., 1997

Oral cancer among dry-cleaning workers

 

 

Possible exposure to tetrachloroethylene

7

1.2 (0.3–4.6)

 

Probable exposure to tetrachloroethylene

4

1.5 (0.2–9.5)

 

Cumulative exposure to tetrachloroethylene (ppm-years)

 

 

1–29 ppm-years

3

1.0 (0.1–7.0)

 

30+ ppm-years

4

1.4 (0.2–8.7)

 

Duration of employment

 

 

1–9 years

6

1.4 (0.3–5.7)

 

10+ years

1

0.4 (0.0–31.6)

aRisk estimate and 95% CI calculated by the committee using standard methods from the observed and expected numbers presented in the original study.

Epidemiologic Studies of Exposure to Organic Solvents and Nasal Cancer

Few studies with sufficient numbers of cases to assess the relationship between exposure to benzene and oral cancer were available. The cohort study by Yin and colleagues (1996a) of

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

benzene-exposed workers in China showed an increased risk of nasopharyngeal cancer among male workers (relative risk [RR]=2.1, 95% CI=0.7–9.3). The RR was 2.4 (95% CI=0.8–10.5).

A cancer mortality study of shoe-manufacturing workers by Fu and colleagues (1996) analyzed the risk associated with exposure to solvents (found mostly in glues) and leather dusts in two cohorts of shoemakers in England and Florence, Italy. Risk of nasal cancer from occupational exposures was strongly increased (SMR=7.41, 95% CI=3.83–12.94) in the English cohort (only one death from nasal cancer was found in the Florence cohort). Exposure was assessed by using job titles, and no specific exposures were identified. It is not clear whether solvents, leather dust, or other agents contributed to the increased risk of nasal cancer.

Hardell and colleagues (1982) conducted a case-control study of nasal and nasopharyngeal cancers and exposure to various agents, including solvents. Although they provided no estimates of relative risk, the committee calculated relative risks and CIs from the raw data provided and found weak associations with exposure to high-grade organic solvents (nasal: OR=1.24, 95% CI=0.51–2.91; nasopharyngeal: OR=1.27, 95% CI=0.47–3.68).

Summary and Conclusion

The only study on exposure to benzene and risk of nasal cancer had a highly imprecise estimate of effect. Other studies are needed to determine whether an association exists. For exposure to solvent mixtures, the English shoemaker study showed a strong association. However, the Swedish case-control study (Hardell et al., 1982) did not corroborate those findings. Table 6.4 identifies the studies reviewed by the committee on nasal cancer. Unless indicated in the table, the study populations include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to solvents under review and nasal cancer.

TABLE 6.4 Selected Epidemiologic Studies—Nasal Cancer and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Specific and Unspecified Mixtures of Organic Solvents

Cohort Studies—Mortality

Yin et al., 1996a

Chinese workers exposed to benzene

 

 

Male

12

2.1 (0.7–9.3)

 

Total

14

2.4 (0.8–10.5)

Fu et al., 1996

Shoemakers in England and Florence

 

 

English cohort

12

7.41 (3.83–12.94)

 

Florence cohort

1

9.09 (0.23–50.65)

Alderson and Rattan, 1980

Male British workers at an isopropyl alcohol plant

1

50.0 (1.3–278.5)a

Case-Control Study

Hardell et al., 1982

Male cases exposed to high-grade organic solvents

 

 

Nasal

8

1.24 (0.51–2.91)a

 

Nasopharyngeal

5

1.27 (0.47–3.68)a

aRisk estimate and 95% CI calculated by the committee using standard methods from the observed and expected numbers presented in the original study.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Epidemiologic Studies of Exposure to Organic Solvents and Laryngeal Cancer

In a cohort of dry cleaners, Blair and colleagues (1990) found 60% excess mortality from laryngeal cancers (SMR=1.6, 95% CI=0.3–4.7). In a case-control study of laryngeal cancer, Vaughan and colleagues (1997) found an association with possible exposure to tetrachloroethylene among dry cleaners (OR=2.3; 95% CI=0.5–10.2) but not among those probably exposed (OR=0.9, 95% CI=0.1–12.9). Risk increased with duration of employment in the dry-cleaning industry but not with increasing cumulative exposure.

Cohorts of workers in ethanol and isopropanol production (Teta et al., 1992) and shoe manufacture (Walker et al., 1993) were evaluated for their cancer mortality, and there was little evidence of an association (total of three exposed deaths).

Summary and Conclusion

For exposure to tetrachloroethylene and dry-cleaning solvents and risk of laryngeal cancer, both studies’ findings were based on very few exposed cases, and this resulted in imprecise estimates of relative risk. Similarly, the studies on solvent mixtures were limited by the small number of exposed cases and lack of positive findings. As a result, there was insufficient evidence to conclude that there were associations between exposure to specific solvents or solvent mixtures and laryngeal cancer. The studies reviewed by the committee are identified in Table 6.5. Unless indicated in the table, the study populations include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to solvents under review and laryngeal cancer.

TABLE 6.5 Selected Epidemiologic Studies—Laryngeal Cancer and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Specific and Unspecified Mixtures of Organic Solvents

Cohort Studies—Mortality

Walker et al., 1993

Ohio shoe manufacturing employees

 

 

Females

2

3.34 (0.40–12.09)

Teta et al., 1992

Male workers at ethanol/isopropanol production plants

 

 

South Charleston

1

1.4 (0.0–8.0)a

 

Texas City

1

3.3 (0.1–18.6)a

Blair et al., 1990

St. Louis, MO, dry-cleaning workers

3

1.6 (0.3–4.7)

Case-Control Study

Vaughan et al., 1997

Laryngeal cancer among dry-cleaning workers

 

 

Possible exposure to tetrachloroethylene

4

2.3 (0.5–10.2)

 

Probable exposure to tetrachloroethylene

1

0.9 (0.1–12.9)

 

Cumulative exposure to tetrachloroethylene (ppm-years)

 

 

1–29 ppm-years

2

2.0 (0.2–17.9)

 

30+ ppm-years

2

2.5 (0.3–19.1)

 

Duration of employment

 

 

1–9 years

3

1.9 (0.3–10.2)

 

10+ years

2

5.5 (0.4–75.0)

aRisk estimate and 95% CI calculated by the committee with standard methods from the observed and expected numbers presented in the original study.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

GASTROINTESTINAL TRACT TUMORS

Description of Case-Control Studies

Several case-control studies were used to evaluate the risks of cancer at gastrointestinal sites in relation to occupational exposures (Table 6.6), and one study was used to assess the risk of colorectal and pancreatic cancers posed by exposure to tetrachloroethylene in drinking water (Paulu et al., 1999).

All the studies but one used interviews with subjects to assess occupational history and in some cases occupational exposures; one study used self-administered questionnaires of next of kin (Kauppinen et al., 1995). The response rates in the latter study were 50% or less, so the results were unlikely to be representative of the entire study population. There were four reports (Dumas et al., 2000; Gérin et al., 1998; Goldberg et al., 2001; Parent et al., 2000) of one multisite case-control study conducted in Montreal in the 1980s. The novel features of the study include use of a mixed control population (cancer and population controls), in-depth interviews to obtain details of each job of each subject, translation of the interviews by a team of industrial hygienists and chemists into semiquantitative indexes of exposure to about 300 physical and chemical agents, and good information on potential confounding factors. Ekstrom and colleagues (1999) investigated gastric cancer and also used experts to attribute exposure on the basis of questionnaires. The study by Paulu and colleagues (1999) of colorectal and pancreatic cancers used estimates of exposure to tetrachloroethylene in drinking water.

Alcohol is a risk factor for esophageal cancer, and this requires consideration in evaluating the association between solvent exposure and esophageal cancer. Two studies (Parent et al., 2000; Vaughan et al., 1997) considered this confounding variable, and one did not (Gérin et al., 1998). Risk factors for other gastrointestinal cancers are less well defined.

Epidemiologic Studies of Exposure to Organic Solvents and Esophageal Cancer

The risk of esophageal cancer was increased in Danish workers who were biologically monitored for a urinary metabolite of trichloroethylene (standardized incidence ratio [SIR]=4.2, 95% CI=1.5–9.2) (Hansen et al., 2001). There was no gradient in risk with cumulative exposure although fairly high relative risks were found for long duration of employment (SIR=6.6, 95% CI=1.8–17). An equivalent risk estimate and 95% CI were also found among workers with high cumulative exposure. Blair and colleagues (1998) reported an excess risk of esophageal cancer mortality among white male aircraft-maintenance workers exposed to trichloroethylene (SMR=5.6, 95% CI=0.7–44.5). In a cohort of aircraft-manufacturing workers in California, Boice and colleagues (1999) found no association between esophageal cancer and potential exposure to trichloroethylene (SMR=0.83, 95% CI=0.34–1.72).

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.6 Description of Case-Control Studies of Gastrointestinal Tract Tumors and Exposure to Organic Solvents

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Canadian studies

Gérin et al., 1998

Male cases and controls, age 35–70 years, diagnosed in 19 large Montreal-area hospitals in 1979–1985 and histologically confirmed for one of 19 anatomic cancer sites; frequency matched by approximate age, population-based controls were also chosen from electoral lists and with random-digit dialing (see also Dumas et al., 2000; Goldberg et al., 2001; Parent et al., 2000)

99 esophageal

251 stomach

497 colon

257 rectal

116 pancreas

1,066 subjects for each site, consisting of 533 population controls and 533 randomly selected subjects from the eligible cancer control group

Benzene

Toluene

Xylene

In-person interviews with specific questions on details of each job subject had; analyzed and coded by a team of chemists and industrial hygienists (about 300 exposures) on semiquantitative scales

Unconditional logistic regression

Age, family income, ethnicity, cigarette smoking, respondent status

Response rates: 82% of all cases, 71% of population controls

Dumas et al., 2000

Same as above (see also Gérin et al., 1998; Goldberg et al., 2001; Parent et al., 2000)

257 rectal

1,295 cancer

533 population

Toluene

Xylene

Methylene chloride

Trichloroethylene

Acetone

See above

See above

Age, education respondent status, cigarette smoking, beer consumption body mass index

Response rates: 84.5% of cases, 72% of population

 

 

Parent et al., 2000

Same as above (see also Dumas et al., 2000; Gérin et al., 1998; Goldberg et al., 2001)

99 esophageal

2,299 cancer

533 poulation

Toluene

Solvents

See above

See above

Age, respondent status, birthplace, educational level, beer consumption, spirits consumption, β-carotene index, cigarette smoking (length and pattern)

Response rates: 75% of cases, 71% of population controls

 

Goldberg et al., 2001

Same as above (see also Dumas et al., 2000; Gérin et al., 1998; Parent et al., 2000)

497 colon

1,514 cancer

533 population

Benzene

Xylene

Toluene

See above

See above

Age, respondent status, ethnicity, nonoccupational factors (such as cigarette smoking, alcohol consumption)

Response rates: 82% of cases, 72% of population controls

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

US studies

Vaughan et al., 1997

Cases, age 20–74 years when diagnosed, identified from a cancer-surveillance system covering 13 counties in western Washington; cases lived in one of the three largest counties and were diagnosed in 1983–1990; population-based controls, frequency-matched by age and sex, identified through RDD

404 esophageal and gastric cardia

724

Tetrachloroethylene

Dry-cleaning work

In-person interviews on occupational history (job titles; including duration, exposure probability, cumulative exposure calculations)

Conditional logistic regression

Age, sex, education, study period, alcohol consumption, cigarette smoking, race

Response rates: 82.9% of cases, 80.3% of controls

Paulu et al., 1999

Cases reported to the Massachusetts Cancer Registry, diagnosed in 1983–1986 among residents of five upper Cape Cod towns; living controls were selected from the records of HCFA and through RDD; deceased controls identified by the state Department of Vital Statistics and Research files

311 colon-rectum

36 pancreas

1,158 (colon-rectum)

622 (pancreas)

Tetrachloroethylene

Calculated relative delivered dose accounting for location and years of residence, water flow, pipe characteristics

Multiple logistic regression

Age at diagnosis, vital status, sex, occupational exposure to solvents; specific cancer risk factors controlled in respective analyses

Response rates: 79% of cases, 76% of HCFA controls, 74% of RDD controls, 79% of next of kin of deceased controls

European studies

Fredriksson et al., 1989

Cases age 30–75 years identified through the Swedish Cancer Registry among patients diagnosed in 1980–1983; cases residents of the Umea region and alive during the study’s data collection; randomly selected population controls from the National Population Register were frequency-matched on age and sex

329 colon

658

Trichloroethylene Organic solvents

Dry-cleaning work Painter

Mailed questionnaire assessing occupational history (job titles); telephone interviews followed if necessary; solvent exposures independently coded by two physicians and one hygienist

Mantel-Haenszel

Age, sex, physical activity

Kauppinen et al., 1995

Deceased cases as of April 1990, age 40–74 years at diagnosis in 1984–1987; identified cases and controls from the Finnish Cancer Registry; controls of similar age and period of diagnosis selected from deceased cases of stomach, colon, or rectal cancer

595 pancreatic

1,622

Solvents

Mailed questionnaire to next of kin assessing lifetime work history (job titles); assignment of exposures by industrial hygienist and use of a job-exposure matrix

Unconditional logistic regression

Age, sex, tobacco smoking, diabetes mellitus, alcohol consumption

Response rates: 47% of cases, 50% of controls

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Ekstrom et al., 1999

Cases, age 40–79 years, residing in one of the five counties, born in Sweden, and diagnosed in 1989–1995, identified and histologically confirmed by participating clinicians from all hospitals in the study area; control subjects randomly selected from the population register

565 gastric

1,164

Organic solvents

In-person interview with professional interviewer; occupational epidemiologists to assess type of exposure and duration from self-reports of exposure and job titles

Unconditional logistic regression

Age, sex

Response rates: 62.4% of cases, 75.9% of controls

Kaerlev et al., 2000

Cases, age 35–69 years, in 10 European countries in 1995–1997 identified from hospital and pathology departments and regional and national cancer registers and histologically confirmed; population-based controls randomly selected from each study base except in Spain, where hospital-based colon cancer cases were used as controls

79 small bowel adenocarcinoma

579 colon cancer

2070 population

Dry-cleaning work

Standard questionnaire administered in person or over the telephone to determine occupational exposures (job or industry titles, specific work tasks), lifestyle factors; occupation and industry codes used to categorize exposure

Unconditional logistic regression

Sex, country, year of birth

Response rates: 74% of cases, 64% of controls

Chinese study

Ji et al., 1999

Cases, age 30–74 years, identified through the Shanghai Cancer Registry among patients to diagnosed from October 1990 to June 1993; cases confirmed through histopathology, gross pathology, or tomography; randomly selected population controls from Shanghai were frequency matched on age and sex

451 pancreatic

1,552

Chemical and rubber work

Rubber work

Printing

In-person interview with professional interviewers to assess occupational history (job titles); job titles coded by the authors using a scheme developed for use in the Third National Census in 1982

Unconditional logistic regression

Age, education, income, cigarette smoking, other occupations

Response rates: 78.2% of cases, 84.5% of controls

NOTE: HCFA=Health Care Financing Administration; RDD=random-digit dialing.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Three cohort studies of workers in the dry-cleaning industry and in aircraft manufacture reported positive associations with esophageal cancer. In a cohort of dry-cleaning union members, Ruder and colleagues (2001) observed associations between esophageal cancer and exposure to tetrachloroethylene (SMR=2.47, 95% CI=1.35–4.14) and between esophageal cancer and long-term exposure to tetrachloroethylene (SMR=5.03, 95% CI=2.41–9.47). The risk of esophageal cancer observed in workers exposed solely to tetrachloroethylene (SMR=2.65, 95% CI=0.85–6.20) was similar to the risk observed in workers exposed to tetrachloroethylene and other dry-cleaning solvents (SMR=2.40, 95% CI=1.10–4.56). Blair and colleagues (1990) found an increased risk of esophageal cancer (SMR=2.1, 95% CI=1.1–3.6) in another cohort of dry-cleaning union workers; the risk among those with high exposure to dry-cleaning solvents was slightly higher (SRR=1.3) than the risk in the referent group with medium exposure. An increased risk associated with potential exposure to tetrachloroethylene was found in a cohort of aircraft-manufacturing workers (SMR=1.47, 95% CI=0.54–3.21) (Boice et al., 1999). However, no exposure-response pattern was apparent.

Vaughan and colleagues (1997) identified cases of several types of cancer, including two morphologic types of esophageal cancer, in examining the risks from occupational exposure. The risk of esophageal squamous cell carcinoma was increased for possible exposure to tetrachloroethylene (OR=3.6, 95% CI=0.5–27.0) and probable exposure (OR=6.4, 95% CI=0.6–68.9). Increases in risk of esophageal adenocarcinoma were found to be associated with possible exposure to tetrachloroethylene but not with probable exposure.

Gérin and colleagues (1998) reported no association between medium or high exposure to benzene and risk of esophageal cancer (OR=0.9, 95% CI=0.3–2.4). In the large cohort of Chinese benzene-exposed factory workers, increased rates of mortality from esophageal cancer were found (RR=1.8, 95% CI=0.8–4.5) (Yin et al., 1996a). The cohort was examined further, including information on cumulative exposure to benzene (Hayes et al., 1996). However, when the cumulative exposure data were categorized, the relative risks did not increase with increasing exposure. The analyses did not adjust for alcohol use, an important risk factor for esophageal cancer and a potential confounder.

Gérin and colleagues (1998) found an increased risk of esophageal cancer with exposure to xylene (OR=1.4, 95% CI=0.5–3.8) or toluene (OR=1.9, 95% CI=0.9–4.2) in the medium or high exposure category. The risk estimates in the low exposure category for those substances were around unity. In a more comprehensive analysis of the same study, Parent and colleagues (2000) found a similar risk associated with “substantial” exposure to toluene (OR=1.5, 95% CI =0.6–3.7). The risk was further increased when the analysis was restricted to cases with squamous cell carcinoma (OR=2.4, 95% CI=0.9–6.4).

The only study of methylene chloride was the comprehensive cohort study of Kodak employees (Hearne and Pifer, 1999; Hearne et al., 1987, 1990). The most recent followup of the cohort included two cases of esophageal cancer in the group exposed to methylene chloride and no excess risk was observed.

A cohort study of workers in US chemical plants evaluated the association between exposure to phenol and esophageal cancer risk (Dosemeci et al., 1991). Although a slightly increased risk was associated with “any” exposure to phenol (SMR=1.6, 95% CI=0.9–2.6), other studies of unspecified mixtures of solvents yielded no increased risks of esophageal cancer (Anttila et al., 1995: SIR=0.41, 95% CI=0.01–2.29; Garabrant et al., 1988: SIR=1.14, 95% CI=0.62–1.92; Parent et al., 2000: OR=1.1, 95% CI=0.7–1.7).

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Summary and Conclusion

Although almost all the studies of esophageal cancer and exposure to tetrachloroethylene and dry-cleaning solvents showed positive associations, the small number of studies (four) and the lack of increased risk with increased exposure led some committee members to favor the inadequate/insufficient category of association. In addition, some committee members expressed concern over the lack of control or adjustment for tobacco and alcohol use (known risk factors for esophageal cancer) (see Chapter 2 and Appendix E for information on smoking as a potential confounder), whereas others believed that the lack of increased risk of lung and bladder cancer in the same studies constituted evidence that confounding alone could not account for the observed increase in esophageal cancer. As a result, several committee members believed that the evidence was inadequate/insufficient to determine whether an association exists between esophageal cancer and exposure to tetrachloroethylene or dry-cleaning solvents, and others felt that the evidence was limited/suggestive of an association. After extensive discussion and deliberation, the committee decided that it could not reach a consensus on the association. Future committees may re-examine this literature and any new studies that are conducted in the interim to clarify the association between exposure to tetrachloroethylene or dry-cleaning solvents and the risk of esophageal cancer.

In cohort studies of workers exposed to trichloroethylene, most risk estimates for esophageal cancer were increased but highly variable (because there were few exposed subjects), and the estimates of risk were not adjusted for known risk factors, including alcohol consumption. For exposure to benzene, xylene, toluene, methylene chloride, phenol, and solvent mixtures, some of the studies provided positive findings while others did not. Most were not statistically precise. The key studies reviewed by the committee on esophageal cancer are identified in Table 6.7. Unless indicated in the table, the study populations include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to solvents under review, other than tetrachloroethylene and dry-cleanings solvents, and esophageal cancer.

TABLE 6.7 Selected Epidemiologic Studies—Esophageal Cancer and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Study—Incidence

Hansen et al., 2001

Biologically monitored Danish workers

 

 

Any exposure

6

4.2 (1.5–9.2)

 

>75 months

4

6.6 (1.8–17)

 

Low cumulative exposure

3

6.5 (1.3–19)

 

High cumulative exposure

3

4.2 (1.5–9.2)

 

Low mean exposure

5

8.0 (2.6–19)

 

High mean exposure

1

1.3 (0.02–7.0)

 

Low employment duration

2

4.4 (0.5–16)

 

High employment duration

4

6.6 (1.8–17)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California, potential routine exposure

7

0.83 (0.34–1.72)

Blair et al., 1998

Aircraft-maintenance workers in Utah, ever exposed

10

5.6 (0.7–44.5)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Tetrachloroethylene and Dry-cleaning Solvents

Cohort Studies—Mortality

Ruder et al., 2001

Dry-cleaning union workers

14

2.47 (1.35–4.14)

 

Long-term exposurea

10

5.03 (2.41–9.47)

 

Tetrachloroethylene-only

5

2.65 (0.85–6.20)

 

Tetrachloroethylene-plus other solvents

9

2.40 (1.10–4.56)

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Potential routine exposure

6

1.47 (0.54–3.21)

 

≥5 years routine or intermittent

3

0.91 (0.13–1.60)

Blair et al., 1990

Dry-cleaning union members in St. Louis, MO

 

 

Dry-cleaning solvents

13

2.1 (1.1–3.6)b

 

Medium exposure (white males)

1

2.9 (0.1–18.6)b

 

Medium exposure (black males)

7

3.6 (1.5–7.6)b

 

High exposure (white males)

0

 

High exposure (black males)

1

2.0 (0.1–11.1)b

Case-Control Study

Vaughan et al., 1997

Cases from the dry-cleaning industry

 

Squamous cell (possible exposure)

2

3.6 (0.5–27.0)

 

Squamous cell (probable)

2

6.4 (0.6–68.9)

 

Adenocarcinoma (possible)

2

1.1 (0.2–5.7)

 

Adenocarcinoma (probable)

1

0.9 (0.1–10.0)

Benzene

Cohort Studies—Mortality

Gérin et al., 1998

Male residents of Montreal, Canada

 

 

Medium or high exposure

5

0.9 (0.3–2.4)

Yin et al., 1996a

Chinese factory workers ever exposed to benzene

27

1.8 (0.8–4.5)

 

Males

25

2.0 (0.9–5.4)

 

Females

2

0.8 (0.1–16.7)

Hayes et al., 1996

Chinese factory workers (cumulative exposure to benzene)

 

 

None

7

1.0

 

<10 ppm-years

5

3.5

 

10–39 ppm-years

1

0.5

 

40–99 ppm-years

3

1.3

 

100–400 ppm-years

5

1.1

 

>400 ppm-years

13

3.2

 

 

 

p-trend=0.09

Xylene and Toluene

Case-Control Studies

Parent et al., 2000

Male residents of Montreal, Canada

 

 

Esophageal

 

 

Any toluene

16

1.2 (0.7–2.2)

 

Substantial toluene

7

1.5 (0.6–3.7)

 

Squamous cell

 

 

Any toluene

15

2.0 (1.0–3.9)

 

Substantial toluene

6

2.4 (0.9–6.4)

Gérin et al., 1998

Male residents of Montreal, Canada

 

 

Medium or high exposure, xylene

5

1.4 (0.5–3.8)

 

Medium or high exposure, toluene

9

1.9 (0.9–4.2)

Methylene Chloride

Cohort Study—Mortality

Hearne and Pifer, 1999

Male Kodak workers in New York state, employed >1 year

 

Methylene chloride cohort

2

0.63 (0.07–2.28)

 

Roll-coating division (external control)

4

1.42 (0.38–3.65)

 

Roll-coating division (internal control)

4

1.40 (0.38–3.58)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Phenol

Cohort Study—Mortality

Dosemeci et al., 1991

Male workers in five US chemical plants

 

Any exposure

15

1.6 (0.9–2.6)

 

No exposure

4

1.0

 

Low exposure

11

0.9

 

Medium exposure

10

2.3

 

High exposure

1

1.1

Unspecified Mixtures of Organic Solvents

Cohort Studies—Incidence

Anttila et al., 1995

Finnish workers biologically monitored for exposure to halogenated hydrocarbons (trichloroethylene, tetrachloroethylene, 1,1,1-trichloroethane)

1

0.41 (0.01–2.29)

Garabrant et al., 1988

Aircraft-manufacturing workers in California, employed >4 years

14

1.14 (0.62–1.92)

Case-Control Study

Parent et al., 2000

Male residents of Montreal, Canada

 

 

Esophageal

39

1.1 (0.7–1.7)

 

Squamous cell

30

1.4 (0.8–2.5)

aLong-term exposure designates 5 years or more of exposure with at least 20 years of latency time.

b95% CI calculated by the committee with standard methods from the observed and expected numbers presented iu the original study.

Epidemiologic Studies of Exposure to Organic Solvents and Stomach Cancer

The risk of stomach cancer was not increased in a cohort of Danish workers biologically monitored for a metabolite of trichloroethylene (Hansen et al., 2001). Another Scandinavian cohort, also of workers biologically monitored for trichloroethylene metabolites, showed an increased risk of stomach cancer (SIR=1.28, 95% CI=0.75–2.04), but higher concentrations of the metabolite were not associated with greater risks (Anttila et al., 1995). Blair and colleagues (1998) reported increased stomach cancer incidence with exposure greater than 5 years (SIR=3.1, 95% CI=0.8–12.1 for 5–25 unit-years of exposure), but the risks did not increase with increasing unit-years of exposure. No increased risk of stomach cancer mortality was reported in the study (SMR=0.9, 95% CI=0.4–1.9). A cohort study of aircraft-manufacturing workers in California had increased mortality (SMR=1.32, 95% CI=0.77–2.12), but an exposure-response analysis was not presented (Boice et al., 1999). A nested case-control study of rubber workers in Ohio showed no increased risks (Wilcosky et al., 1984).

Two cohort studies of workers in the dry-cleaning industry suggested no association between occupational dry-cleaning solvent exposure and stomach cancer (Blair et al., 1990; Ruder et al., 1994). Boice and co-workers (1999) found an increased risk of stomach cancer in the California cohort of aircraft- manufacturing workers (SMR=1.42, 95% CI=0.57–2.93).

No associations were found in the large cohort of Chinese benzene-exposed workers (Hayes et al., 1996; Yin et al., 1996b). In the Montreal case-control study (Gérin et al., 1998), the risk of stomach cancer was associated with medium exposure to benzene (OR=1.5, 95% CI =0.8–3.2) and high exposure (OR=1.3, 95% CI=0.5–3.2). A slightly increased risk was seen in rubber workers potentially exposed to benzene (Wilcosky et al., 1984).

In a cohort study of rotogravure workers exposed primarily to toluene, Svensson and colleagues (1990) reported an increased risk of stomach cancer mortality (SMR=2.72, 95% CI

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

=1.09–5.61). Similar risk estimates were reported among subjects with 5 years of exposure or more and at least a 10-year latency period. The cancer incidence results were similar. In the Montreal case-control study (Gérin et al., 1998), the risk of stomach cancer was increased with high exposure to toluene (OR=1.7, 95% CI=0.6–4.8), and a similar risk estimate was reported for exposure to xylene (OR=1.8, 95% CI=0.3–9.5).

In the Finnish study of workers biologically monitored for aromatic hydrocarbon exposures (xylene, toluene, and styrene), Anttila and colleagues (1998) found increased stomach cancer incidence (SIR=1.18, 95% CI=0.54–2.23). In a study of rubber workers potentially exposed to xylene or toluene, no excess risks of stomach cancer were found (Wilcosky et al., 1984).

Two cohort studies of workers exposed to methylene chloride found no persuasive evidence of associations between stomach cancer and exposure (Hearne and Pifer, 1999; Tomenson et al., 1997). Stomach cancer risk was somewhat increased in the Kodak cohort (Hearne and Pifer, 1999), but Tomenson and co-workers (1997) found no association in the cohort of cellulose triacetate workers they followed.

Dosemeci and colleagues (1991) examined stomach cancer associated with exposure to phenol in a cohort of workers at five US chemical plants. No increased risk was found to be associated with “any” exposure (SMR=0.8, 95% CI=0.5–1.3) or with categories of increasing exposure. The case-control study conducted among rubber workers included results for several specific solvents (naphthas, ethanol, acetone, isopropanol, and toluene mixture) and stomach cancer risk (Table 6.8). Risk was slightly increased with some exposures but not others (Wilcosky et al., 1984).

In several cohort studies (Acquavella et al., 1993; Anttila et al., 1995; Berlin et al., 1995; Costantini et al., 1989; Fu et al., 1996; Garabrant et al., 1988) and one case-control study (Ekstrom et al., 1999), the association between stomach cancer and exposure to mixed solvents was examined. Except for the Florence cohort of shoemakers (Fu et al., 1996), the studies showed no association with stomach cancer risk. Fu and colleagues (1996) reported a 92% excess risk of stomach cancer associated with solvents used by shoemakers.

Summary and Conclusion

With only one study showing a highly variable positive association, the committee concluded that the data were insufficient to determine whether an association exists between the risk of stomach cancer and exposure to trichloroethylene. For exposure to tetrachloroethylene and dry-cleaning solvents, benzene, xylene, toluene, methylene chloride, phenol, other specific solvents, and solvent mixtures, the results were mixed. Table 6.8 identifies the data points considered by the committee in making its conclusion regarding association for stomach cancer. Unless indicated in the table, the populations cited include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to solvents under review and stomach cancer.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.8 Selected Epidemiologic Studies—Stomach Cancer and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Studies—Incidence

Hansen et al., 2001

Biologically monitored Danish workers

3

0.8 (0.2–2.3)

Blair et al., 1998

Aircraft-maintenance workers in Utah

 

 

No trichloroethylene exposure

6

1.5 (0.4–6.0)

 

<5 unit-years

1

0.3 (0.1–2.6)

 

5–25 unit-years

7

3.1 (0.8–12.1)

 

>25 unit-years

6

2.0 (0.5–8.1)

Anttila et al., 1995

Finnish workers biologically monitored for exposure

 

 

Years since first measurement

 

 

0–9 years

6

1.32 (0.48–2.87)

 

10–19 years

4

0.63 (0.17–1.60)

 

20+ years

7

2.98 (1.20–6.13)

 

Whole period

17

1.28 (0.75–2.04)

 

Mean personal U-TCA level

 

 

<100 µmol/L

12

1.65 (0.98–1.39)

 

100+ µmol/L

4

0.91 (0.25–2.32)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California, potential routine exposure

17

1.32 (0.77–2.12)

Blair et al., 1998

Aircraft-maintenance workers in Utah, employed >1 year

23

0.9 (0.4–1.9)

Wilcosky et al., 1984

Male rubber workers in Ohio, exposed >1 year

5

1.0

Tetrachloroethylene and Dry-cleaning Solvents

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California, potential routine exposure

7

1.42 (0.57–2.93)

Ruder et al., 1994

Dry-cleaning labor-union workers

5

0.61 (0.20–1.43)

 

Males

2

0.43 (0.05–1.54)

 

Females

3

0.86 (0.18–2.53)

Blair et al., 1990

Dry-cleaning union members in St. Louis, MO

 

 

Dry-cleaning solvents

11

0.8 (0.4–1.4)

Benzene

Cohort Studies—Incidence

Yin et al., 1996a

Chinese factory workers ever exposed to benzene

85

0.9 (0.7–1.4)

 

Males

71

0.9 (0.6–1.4)

 

Females

14

1.0 (0.4–2.8)

Hayes et al., 1996

Chinese factory workers (cumulative exposure to benzene)

 

 

None

43

1.0

 

<10 ppm-years

6

0.6

 

10–39 ppm-years

13

1.0

 

40–99 ppm-years

12

0.9

 

100–400 ppm-years

25

1.0

 

400+ ppm-years

27

1.2

 

 

 

p-trend=0.63

Case-Control Study

Gérin et al., 1998

Male residents of Montreal, Canada

 

 

Benzene, medium exposure

11

1.5 (0.8–3.2)

 

Benzene, high exposure

6

1.3 (0.5–3.2)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Ris (95% CI)

Xylene and Toluene

Cohort Studies—Incidence

Anttila et al., 1998

Finnish workers biologically monitored for exposure to aromatic hydrocarbons (styrene, toluene, xylene)

9

1.18 (0.54–2.23)

Svensson et al., 1990

Male rotogravure workers in Sweden

7

2.34 (0.94–4.82)

≥5 years exposure with ≥10 years latency (primarily to toluene)

5

2.18 (0.71–5.09)

Cohort Studies—Mortality

Svensson et al., 1990

Male rotogravure workers in Sweden

7

2.72 (1.09–5.61)

≥5 years exposure with ≥10 years latency (primarily to toluene)

5

2.53 (0.82–5.91)

Wilcosky et al., 1984

Male rubber workers in Ohio, exposed >1 year

 

Xylenes

3

0.53

 

Toluene

1

 

Benzene

12

1.3

Case-Control Study

Gérin et al., 1998

Male residents of Montreal, Canada

 

 

Xylene, medium exposure

7

1.0 (0.4–2.3)

 

Xylene, high exposure

2

1.8 (0.3–9.5)

 

Toluene, medium exposure

7

1.0 (0.4–2.2)

 

Toluene, high exposure

5

1.7 (0.6–4.8)

Methylene Chloride

Cohort Studies—Mortality

Hearne and Pifer, 1999

Male Kodak workers in New York state, employed >1 year

 

Methylene chloride cohort

6

1.40 (0.51–3.04)

 

Roll-coating division (external control)

5

1.25 (0.40–2.91)

 

Roll-coating division (internal control)

5

1.26 (0.41–2.93)

Tomenson et al., 1997

Male cellulose triacetate film workers in Brantham, UK, ever employed

6

0.63 (0.23–1.37)

Other Organic Solvents

Cohort Studies—Mortality

Dosemeci et al., 1991

Male workers in five US chemical plants

 

Phenol

 

 

No exposure

10

1.1

 

Any exposure

18

0.8 (0.5–1.3)

 

Low exposure

11

1.0

 

Medium exposure

5

0.5

 

High exposure

2

1.1

Wilcosky et al., 1984

Male rubber workers in Ohio, exposed >1 year

 

Specialty naphthas

18

1.1

 

Ethanol

8

1.1

 

Acetone

1

 

Isopropanol

14

1.4

 

Phenol

6

1.4

 

VM&P naphtha

3

1.1

 

Solvent “A” (toluene mixture)

15

1.4

Unspecified Mixtures of Organic Solvents

Cohort Studies—Incidence

Anttila et al., 1995

Finnish workers biologically monitored for exposure to halogenated hydrocarbons (trichloroethylene, tetrachloroethylene, 1,1,1-trichloroethane)

19

1.28 (0.77–1.99)

Berlin et al., 1995

Swedish patients with acute solvent-related disorders

2

0.5 (0.1–1.9)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Cohort Studies—Mortality

Fu et al., 1996

Shoemakers in England and Florence

 

 

English cohort

 

 

Probable solvents

29

0.88 (0.59–1.27)

 

High solvent

5

1.72 (0.56–4.02)

1.92 (1.02–3.29)

 

Florence cohort

 

1.93 (1.00–3.36)

 

Probable solvents

13

 

 

High solvent

12

 

Acquavella et al., 1993

Metal-components manufacturing workers

 

Solvents, ever exposed

0

0.0 (0.0–17.6)

Costantini et al., 1989

Male leather workers in Tuscany, Italy, employed >6 months

6

0.43 (0.16–0.94)

Garabrant et al., 1988

Aircraft manufacturing workers in California, employed >4 months

9

0.40 (0.18–0.76)

Case-Control Study

Ekstrom et al., 1999

Residents of two regions in Sweden

 

Organic solvents, ever exposed

232

1.08 (0.87–1.33)

NOTE: U-TCA=urinary metabolite of trichloroethylene.

Epidemiologic Studies of Exposure to Organic Solvents and Colon Cancer

In a study conducted in Sweden (Fredriksson et al., 1989), a 640% excess risk of colon cancer (OR=7.4, 95% CI=1.1–47.0) was found with exposure to trichloroethylene among dry cleaners. Two cohort studies of aircraft workers did not show increased mortality from colon cancer (Blair et al., 1998; Boice et al., 1999). However, the incidence of colon cancer was increased in the cohort of aircraft-maintenance workers in Utah (Blair et al., 1998). Incidence increased with increasing unit-years of exposure (Table 6.9). Colon cancer risk was not increased in two cohorts biologically monitored for exposure to trichloroethylene (Anttila et al., 1995; Hansen et al., 2001). Anttila and co-workers found no increased risk associated with number of years since first measurement, which represents an approach to account for latency.

A case-control study of colon cancer in Sweden conducted by Fredriksson and colleagues (1989) reported an increased risk of colon cancer among female dry cleaners (OR=2.0, 95% CI=0.5–7.1). Paulu and colleagues (1999) conducted a case-control study of residents of Cape Cod that showed an increased risk of colorectal cancers with tetrachloroethylene exposure in drinking water (OR11 year latency=2.0, 95% CI=0.6–5.8) for exposures less than the 50th percentile, whereas the risk was somewhat lower when exposure equal to or greater than the 50th percentile was considered (OR11 year latency=1.5, 95% CI=0.4–4.4). A cohort of dry-cleaning workers experienced an increased risk of intestinal (excluding rectal) cancer (SMR=1.48, 95% CI=1.01–2.09), and findings were similar in a subcohort exposed to both tetrachloroethylene and other solvents; however, the risk was not similarly increased among those exposed only to tetrachloroethylene (Ruder et al., 2001). Because the analysis was based on cases of “intestinal” cancer, the findings are difficult to interpret; different classes of intestinal tract cancers, including cancers of the small and large intestines, are derived from distinct cells and may have different etiology. Real effects may be masked when diseases with different

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

etiology are investigated as one disease. Furthermore, when exposure was restricted to tetrachloroethylene, no increase risk was apparent.

Goldberg and colleagues (2001) reported an increased colon cancer risk with exposure to benzene at “substantial,” “medium,” and “high” levels; and risks increased by about 10% for every 10 years of exposure. In the large cohort of Chinese benzene-exposed workers, no increased risks of colorectal cancer were found (RR=0.9, 95% CI=0.5–1.7) (Yin et al., 1996a). Increasing cumulative exposure did not appear to be associated with increased risk (Hayes et al., 1996).

In detailed followup analyses of the population-based case-control study of occupational exposure and cancer in Montreal, the risk of colon cancer was increased with “substantial” exposure to xylene (OR=1.5, 95% CI=0.6–3.7) and toluene (OR=1.5, 95% CI=0.8–2.5) (Goldberg et al., 2001). Risk estimates increased with increasing concentration of xylene and toluene, and risk increased by 20% for every 10 years of exposure. Levels were assessed on the basis of responses to interviews and structured questionnaires that were coded by a team of chemists and industrial hygienists. Low levels were assumed if a person had been exposed peripherally or at normal levels, and high levels were assumed if a person directly handled a product that contained one of the chemicals of concern. Medium concentration fell between those two.

Colon cancer risk was not increased among workers monitored for hydrocarbons, which included styrene, toluene, and xylene (Anttila et al., 1998). In a cohort study of cancer incidence and mortality in toluene-exposed rotogravure workers, Svensson and colleagues (1990) reported an increased risk of combined colon and rectal cancer mortality (SMR=2.18, 95% CI=0.88–4.49) and incidence (SIR=1.49, 95% CI=0.68–2.84). Similar risk estimates were reported among subjects with prolonged exposure.

Two cohort studies of workers exposed to methylene chloride reported no increased colon cancer risk in the roll-coating division at Kodak (Hearne and Pifer, 1999) or in cellulose triacetate film workers in the UK (Tomenson et al., 1997). The committee concluded that these results did not indicate excess risk of colon cancer posed by exposure to methylene chloride.

One study reported no increased colon cancer risk associated with exposure to phenol (Dosemeci et al., 1991). Risk of colon cancer was not associated with exposure to solvents in a cohort of patients with solvent-related disorders (Berlin et al., 1995) or among workers biologically monitored for halogenated hydrocarbons (Anttila et al., 1995). Self-reported occupation as a painter and occupational exposure to solvents were each associated with colon cancer in a study conducted in Sweden (Fredriksson et al., 1989).

Summary and Conclusion

Because the increased risk and exposure-response pattern support an association between colon cancer and exposure to trichloroethylene, several committee members believed that the evidence was limited/suggestive of an association. Other members felt that the positive associations were balanced by the negative findings in other cohort studies and in studies with biologic monitoring of metabolites of exposure. Therefore, the committee decided not to have a consensus conclusion. Additional research will help to clarify the relationship between exposure to trichloroethylene and the risk of colon cancer.

Results of the three studies on tetrachloroethylene and dry-cleaning solvents are insufficient to determine whether an association exists for colon cancer, because of the described

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

limitations of the cohort study and because the estimates of risk in the case-control studies are imprecise, being based on few exposed cases.

The body of evidence on colon cancer and exposure to benzene and mixtures of toluene and xylene was rather small (five studies), including one high quality case-control study (Goldberg et al., 2001) and two cohort studies that included exposures other than benzene, toluene, or xylene in their analyses (Anttila et al., 1998; Svensson et al., 1990). Anttila and colleagues (1998) assessed the association between colon cancer and hydrocarbons that included styrene in addition to toluene and xylene; and Svensson and colleagues (1990), who examined a cohort of Swedish rotogravure printers, focused primarily on toluene but acknowledged the presence of other solvents and chemical agents. Although one study showed an association with exposure to benzene, toluene, and xylene (Goldberg et al., 2001), the large Chinese factory-worker study (Hayes et al., 1996; Yin et al., 1996b), which combined colon and rectal cancer, did not. The strengths of the Goldberg and colleagues study (2001) included adjustment for most known risk factors and occupational exposures; use of incident, histologically-confirmed cases; and an independent assessment of exposure by experts. The strength of the Chinese cohort study (Hayes et al., 1996) was the relatively accurate estimates of exposure, but its limitations included use of mortality instead of incidence and lack of assessment of confounding factors. As a result, some committee members concluded that the evidence was limited/suggestive of an association, and others concluded that it was insufficient to determine whether an association exists. After much deliberation, the committee decided that it could not reach a consensus on an association of colon cancer and exposure to benzene and mixtures of toluene and xylene. Further studies on these exposure-outcome relationships may provide evidence as to whether an association exists.

For exposure to methylene chloride, phenol, and mixtures of solvents, the studies did not provide any evidence of an association between exposure and risk for colon cancer. All of the studies reviewed by the committee are identified below in Table 6.9, and unless indicated, the study populations include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to solvents under review, other than trichloroethylene, benzene, toluene, and xylene, and colon cancer.

TABLE 6.9 Selected Epidemiologic Studies—Colon Cancer and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylenee

Cohort Studies-Incidence

Hansen et al., 2001

Biologically monitored Danish workers

 

 

Males

5

0.7 (0.2–1.6)

 

Females

1

0.7 (0.01–4.0)

Blair et al., 1998

Aircraft-maintenance workers in Utah

 

 

No trichloroethylene exposure

22

4.1 (1.4–11.8)

 

≤5 unit-years

15

2.9 (1.0–8.9)

 

5–25 unit-years

14

4.3 (1.4–13.0)

 

>25 unit-years

23

5.7 (2.0–16.7)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Ris (95% CI)

Anttila et al., 1995

Biologically monitored Finnish workers

 

 

Years since first measurement:

 

 

0–9 years

3

1.23 (0.25–3.59)

 

10–19 years

3

0.62 (0.13–1.80)

 

20+ years

2

0.92 (0.11–3.31)

 

Whole period

8

0.84 (0.36–1.66)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California, potential routine exposure

30

1.07 (0.72–1.52)

Blair et al., 1998

Aircraft-maintenance workers in Utah

54

1.4 (0.8–2.4)

 

Males

 

 

No trichloroethylene exposure

21

1.5 (0.7–3.3)

 

<5 unit-years

19

1.5 (0.7–3.3)

 

5–25 unit-years

12

1.5 (0.7–3.6)

 

>25 unit-years

15

1.5 (0.7–3.3)

Case-Control Study

Fredriksson et al., 1989

Residents of Sweden

 

Trichloroethylene

NA

1.5 (0.4–5.7)

 

Trichloroethylene among dry cleaners

NA

7.4 (1.1–47.0)

Tetrachloroethylene and Dry-cleaning Solvents

Cohort Study—Mortality

Ruder et al., 2001

Dry-cleaning labor-union workers (intestine)

32

1.48 (1.01–2.09)

 

Long-term exposurea

13

1.48 (0.79–2.58)

 

Tetrachloroethylene only

8

1.18 (0.51–2.33)

Case-Control Studies

Paulu et al., 1999

Residents of Cape Cod, MA

 

 

Colon-rectum (11-year latency)

 

 

≤Median

6

2.0 (0.6–5.8)

 

>Median

5

1.5 (0.4–4.4)

Fredriksson et al., 1989

Residents of Sweden

 

Female dry cleaners

5

2.0 (0.5–7.1)

Benzene

Cohort Studies—Mortality

Yin et al., 1996a

Chinese factory workers ever exposed to benzene (colon-rectum)

 

 

Total

34

0.9 (0.5–1.7)

 

Males

24

1.1 (0.5–2.3)

 

Females

10

0.7 (0.3–2.0)

Hayes et al., 1996

Chinese factory workers (colorectal; benzene cumulative exposure)

 

 

None

17

1.0

 

<10 ppm-years

7

1.5

 

10–39 ppm-years

4

0.7

 

40–99 ppm-years

3

0.5

 

100–400 ppm-years

8

0.8

 

400+ ppm-years

12

1.4

 

 

 

p-trend=0.91

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Case-Control Study

Goldberg et al., 2001

Male residents of Montreal, Canada

 

 

Substantial exposure

21

1.6 (0.9–2.8)

 

Low concentration

39

0.8 (0.5–1.2)

 

Medium concentration

26

1.5 (0.9–2.4)

 

High concentration

6

3.4 (1.0–11.2)

 

Duration (10 years)

71

1.1 (0.9–1.2)

Xylene and Toluene

Cohort Studies—Incidence

Anttila et al., 1998

Finnish workers biologically monitored for exposure to aromatic hydrocarbons (styrene, toluene, and xylene)

2

0.34 (0.04–1.21)

Svensson et al., 1990

Male rotogravure workers in Sweden—colon-rectum

9

1.49 (0.68–2.84)

 

≥5 yrs exposure with >10 yrs latency (primarily toluene)

8

1.74 (0.75–3.43)

Cohort Study—Mortality

Svensson et al., 1990

Male rotogravure workers in Sweden—colon-rectum

7

2.18 (0.88–4.49)

 

≥5 yrs exposure with >10 yrs latency (primarily to toluene)

6

2.41 (0.89–5.25)

Case-Control Study

Goldberg et al., 2001

Male residents of Montreal, Canada

 

 

Xylene

 

 

Substantial exposure

10

1.5 (0.6–3.7)

 

Low concentration

44

1.4 (0.9–2.0)

 

Medium concentration

11

1.7 (0.8–3.5)

 

High concentration

5

4.0 (1.1–15.1)

 

Duration (10 years)

60

1.2 (1.0–1.4)

 

Toluene

 

 

Substantial exposure

27

1.5 (0.8–2.5)

 

Low concentration

31

1.3 (0.8–2.0)

 

Medium concentration

31

1.3 (0.8–2.1)

 

High concentration

9

2.4 (1.0–5.7)

 

Duration (10 years)

71

1.2 (1.0–1.4)

Methylene Chloride

Cohort Studies—Mortality

Hearne and Pifer, 1999

Male Kodak workers in New York state (colon-rectum), employed >1 year

 

 

Methylene chloride cohort

15

1.15 (0.64–1.90)

 

Roll-coating division (external control)

10

0.75 (0.36–1.37)

 

Roll-coating division (internal control)

10

0.87 (0.42–1.60)

Tomenson et al., 1997

Male cellulose triacetate film workers in Brantham, UK, ever employed

6

0.90 (0.33–1.96)

Phenol

Cohort Study—Mortality

Dosemeci et al., 1991

Male workers in five US chemical plants

 

Phenol, any exposure

33

0.9 (0.6–1.3)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Ris (95% CI)

Unspecified Mixtures of Organic Solvents

Cohort Studies—Incidence

Anttila et al., 1995

Finnish workers biologically monitored for exposure to halogenated hydrocarbons (trichloroethylene, tetrachloroethylene, 1,1,1-trichloroethane)

8

0.74 (0.32–1.44)

Berlin et al., 1995

Swedish patients with acute solvent-related disorders

2

0.6 (0.1–2.2)

Case-Control Study

Fredriksson et al., 1989

Residents of Sweden

 

Organic solvents

 

 

High-grade

NA

2.1 (0.8–5.8)

 

Low-grade

NA

1.3 (0.8–2.0)

 

Painters

7

3.0 (0.9–9.2)

NOTE: NA=not available.

aLong-term exposure designates 5 years or more of exposure with at least 20 years of latency.

Epidemiologic Studies of Exposure to Organic Solvents and Rectal Cancer

The two studies of workers biologically monitored for a metabolite of trichloroethylene showed imprecise associations (Anttila et al., 1995; Hansen et al., 2001). Risks did not increase with increasing mean urinary trichloroacetic acid, a metabolite of trichloroethylene (SIR=0.85, 95% CI=0.10–3.07) (Anttila et al., 1995).

No associations with rectal cancer mortality were found in the studies of aircraft and aerospace workers. Blair and colleagues (1998) found no increased risk of rectal cancer among aircraft maintenance workers in Utah (SMR=0.4, 95% CI=0.1–1.5). Among a trichloroethylene-exposed subcohort of aerospace workers, the SMR for “high” exposure was 1.38 (95% CI=0.45–3.21) (Morgan et al., 1998). In the cohort of aircraft-manufacturing workers in California, the SMR was 1.29 (95% CI=0.59–2.45) (Boice et al., 1999). Rectal cancer risk was found to be increased in the Montreal case-control study with “any” exposure to trichloroethylene (OR=2.0, 95% CI=1.0–3.9), but it was much lower with “substantial” exposure to trichloroethylene (OR=0.9, 95% CI=0.3–3.2) (Dumas et al., 2000).

The cohort of dry-cleaning union members experienced an excess risk of rectal cancer mortality after exposure to tetrachloroethylene and other dry-cleaning solvents (SMR=2.16, 95% CI=0.86–4.45) (Ruder et al., 2001). Paulu and colleagues (1999) conducted a case-control study of residents of Cape Cod that showed an increased risk of colorectal cancers with exposure to tetrachloroethylene from drinking water (OR11 year latency=2.0, 95% CI=0.6–5.8) when exposure was less than the 50th percentile; the risk was somewhat lower when exposure was equal to or greater than the 50th percentile (OR11 year latency=1.5, 95% CI=0.4–4.4). Because the analysis was based on cases of “intestinal” cancer, the findings are difficult to interpret; different classes of intestinal tract cancers, including cancers of the small and large intestines, are derived from different cells and may have different etiology. Real effects may be masked when diseases with different etiology are studied as one disease. No cases were found in workers exposed only to tetrachloroethylene.

In the large cohort of Chinese benzene-exposed workers, no increased risk of colorectal cancer was found (RR=0.9, 95% CI=0.5–1.7) (Yin et al., 1996a). Increasing cumulative exposure did not appear to be associated with increased risk (Hayes et al., 1996). Only in the population-based case-control study of occupational exposure and cancer in Montreal was

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

exposure to xylene, toluene, and benzene assessed. Gérin and colleagues (1998) found no association with “high” exposure to benzene (OR=0.8, 95% CI=0.3–2.5). In additional analyses of the data, Dumas and colleagues (2000) found increased rectal cancer risk with “substantial” exposure to xylene (OR=2.9, 95% CI=1.1–7.3) and “substantial” exposure to toluene (OR=1.7, 95% CI=1.0–3.0).

In the cohort study of Finnish workers biologically monitored for exposure to solvents, rectal cancer incidence was associated with exposure to styrene, xylene, and toluene (aromatic hydrocarbons) (SIR=1.88, 95% CI 0.81–3.71) (Anttila et al., 1998).

The cohort study of cellulose triacetate film workers in the UK showed no increased risk of rectal cancer (SMR=0.44, 95% CI=0.05–1.57) (Tomenson et al., 1997). Dumas and colleagues (2000) reported an increased risk of rectal cancer with “any” exposure to methylene chloride (OR=1.2, 95% CI=0.5–2.8); the risk was higher with “substantial” exposure (OR=3.8, 95% CI=1.1–12.9).

The association between rectal cancer and exposure to phenol was assessed in one cohort study (Dosemeci et al., 1991); there was no strong evidence of an association (SMR=1.4, 95% CI=0.8–2.2). Dumas and colleagues (2000) found an increased risk with “any” exposure to acetone (OR=2.3, 95% CI=1.1–4.7), which increased to an OR of 4.8 (95% CI=1.8–13.0) with “substantial” exposure.

There were three cohort studies of rectal cancer and exposure to unspecified mixtures of solvents. An early study of aircraft-manufacturing workers (Garabrant et al., 1988) and a study of Swedish patients with solvent-related disorders (Berlin et al., 1995) showed no associations with rectal cancer (SIR=1.04 and 0.99, respectively). Anttila and colleagues (1995) reported an increased risk of rectal cancer among workers biologically monitored for halogenated hydrocarbon exposure (SIR=1.63, 95% CI=0.87–2.78).

Summary and Conclusion

In summary, there was inconclusive evidence of an association between trichloroethylene and rectal cancer. The main limitation of the studies was the small number of exposed cases; this limits the precision of the estimates and the statistical power to detect associations. For exposure to tetrachloroethylene and dry-cleaning solvents, the committee concluded that the results did not indicate an excess risk of rectal cancer. Despite the suggestive findings on toluene and xylene, only two studies specifically examined rectal cancer risk. To determine whether an association exists, other high-quality studies are required.

Because there were only two studies to draw inferences from, the evidence for methylene chloride was inadequate. Furthermore, the committee could not draw conclusions from single studies of each exposure although there was suggestive evidence from the high-quality study of Dumas and colleagues regarding exposure to acetone. For exposure to solvent mixtures, the findings were mixed with no persuasive evidence of an association.

Table 6.10 identifies the studies and data points considered by the committee in making its conclusion regarding association. Unless indicated, the study populations identified in the table include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to solvents under review and rectal cancer.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.10 Selected Epidemiologic Studies—Rectal Cancer and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Studies—Incidence

Hansen et al., 2001

Biologically monitored Danish workers

7

1.3 (0.5–2.7)

Anttila et al., 1995

Biologically monitored Finnish workers

 

 

Entire period since first measurement

12

1.71 (0.88–2.98)

 

0–9 years

3

1.59 (0.33–4.64)

 

10–19 years

8

2.22 (0.96–4.36)

 

20+ years

1

0.67 (0.02–3.72)

 

Mean personal U-TCA level:

 

 

<100 µmol/L

9

2.34 (1.07–4.44)

 

100+ µmol/L

2

0.85 (0.10–3.07)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California, potential routine exposure

9

1.29 (0.59–2.45)

Blair et al., 1998

Aircraft-maintenance workers in Utah, employed >1 year

5

0.4 (0.1–1.5)

Morgan et al., 1998

Aerospace workers in Arizona

 

Trichloroethylene-exposed subcohort

6

1.06 (0.39–2.31)

 

High trichloroethylene-exposure

5

1.38 (0.45–3.21)

Case-Control Study

Dumas et al., 2000

Male residents of Montreal, Canada

 

 

Any exposure

12

2.0 (1.0–3.9)

 

Substantial exposure

3

0.9 (0.3–3.2)

Tetrachloroethylene and Dry-cleaning Solvents

Cohort Study—Mortality

Ruder et al., 2001

Dry-cleaning labor-union workers (intestine)

 

 

Tetrachloroethylene plus other solvents

7

2.16 (0.86–4.45)

Case-Control Study

Paulu et al., 1999

Residents of Cape Cod, MA

 

 

Colon-rectum (11-year latency)

 

 

≤Median

6

2.0 (0.6–5.8)

 

>Median

5

1.5 (0.4–4.4)

Benzene

Cohort Studies—Mortality

Yin et al., 1996a

Chinese factory workers ever exposed to benzene (colon-rectum)

 

 

Total

34

0.9 (0.5–1.7)

 

Males

24

1.1 (0.5–2.3)

 

Females

10

0.7 (0.3–2.0)

Hayes et al., 1996

Chinese factory workers (colorectal; benzene cumulative exposure)

 

 

None

17

1.0

 

<10 ppm-years

7

1.5

 

10–39 ppm-years

4

0.7

 

40–99 ppm-years

3

0.5

 

100–400 ppm-years

8

0.8

 

400+ ppm-years

12

1.4

 

 

 

p-trend=0.91

Case-Control Study

Gérin et al., 1998

Male residents of Montreal, Canada

 

 

High exposure

4

0.8 (0.3–2.5)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Xylene and Toluene

Cohort Study—Incidence

Anttila et al., 1998

Finnish workers biologically monitored for exposure to aromatic hydrocarbons (styrene, toluene, xylene)

8

1.88 (0.81–3.71)

Case-Control Study

Dumas et al., 2000

Male residents of Montreal, Canada

 

 

Xylene

 

 

Any exposure

39

1.3 (0.9–1.9)

 

Substantial exposure

7

2.9 (1.1–7.3)

 

Toluene

 

 

Any exposure

50

1.4 (1.0–2.0)

 

Substantial exposure

17

1.7 (1.0–3.0)

Methylene Chloride

Cohort Study—Mortality

Tomenson et al., 1997

Male cellulose triacetate film workers in Brantham, UK, ever employed

2

0.44 (0.05–1.57)

Case-Control Study

Dumas et al., 2000

Male residents of Montreal, Canada

 

 

Any exposure

7

1.2 (0.5–2.8)

 

Substantial exposure

5

3.8 (1.1–12.9)

Phenol and Acetone

Cohort Study—Mortality

Dosemeci et al., 1991

Male workers in five US chemical plants

 

Phenol

 

 

No exposure

6

1.1

 

Any exposure

18

1.4 (0.8–2.2)

 

Low exposure

9

1.4

 

Medium exposure

9

1.5

 

High exposure

0

0

Case-Control Study

Dumas et al., 2000

Male residents of Montreal, Canada

 

 

Acetone

 

 

Any exposure

11

2.3 (1.1–4.7)

 

Substantial exposure

8

4.8 (1.8–13.0)

Unspecified Mixtures of Organic Solvents

Cohort Studies—Incidence

Anttila et al., 1995

Finnish workers biologically monitored for exposure to halogenated hydrocarbons (trichloroethylene, tetrachloroethylene, 1,1,1-trichloroethane)

13

1.63 (0.87–2.78)

Berlin et al., 1995

Swedish patients with acute solvent-related disorders

2

0.99 (0.1–3.6)

Cohort Study—Mortality

Garabrant et al., 1988

Aircraft-manufacturing workers in California, employed >4 years

15

1.04 (0.59–1.73)

NOTE: U-TCA=urinary metabolite of trichloroethylene.

Epidemiologic Studies of Exposure to Organic Solvents and Pancreatic Cancer

Several cohort studies of trichloroethylene-exposed workers showed no increased risk of pancreatic cancer. In the cohort of aircraft-manufacturing workers examined by Boice and

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

colleagues (1999), risk was shown to decrease (SMR=0.41, 95% CI=0.17–0.85) with potential exposure to trichloroethylene. Blair and co-workers (1998) also found no association between pancreatic cancer incidence and all categories of “unit-years” of exposure. Pancreatic cancer mortality was weakly increased in the same study (SMR=1.2, 95% CI=0.6–2.3). A third cohort study of aerospace workers in Arizona also found no risk of pancreatic cancer posed by exposure to high levels of trichloroethylene (SMR=0.66, 95% CI=0.24–1.43) (Morgan et al., 1998).

Two cohorts of workers biologically monitored for metabolites of trichloroethylene reported mixed findings. Hansen and colleagues (2001) found no association (SIR=1.0, 95% CI =0.2–3.0) between the trichloroethylene metabolite and pancreatic cancer, whereas Anttila and colleagues (1995) found an increased risk (SIR=1.61, 95% CI=0.81–2.88). However, no exposure-response relationship was indicated as the mean concentration of the metabolite increased.

Ruder and colleagues (2001) found an association between pancreatic cancer and exposure to tetrachloroethylene and other solvents (SMR=1.89, 95% CI=1.06–3.11), but no increase in risk was found in the subcohort of workers exposed only to tetrachloroethylene (SMR =0.80, 95% CI=0.17–2.35). An increased risk of pancreatic cancer was found in the cohort of workers monitored for solvents, including tetrachloroethylene (SIR=3.08, 95% CI=0.63–8.99) (Anttila et al., 1995).

The case-control study of multiple cancer sites performed by Gérin and colleagues (1998) indicated no association between pancreatic cancer and medium or high exposure to toluene (OR=0.6, 95% CI=0.2–2.2), xylene (OR=1.1, 95% CI=0.4–3.3), or benzene (OR=0.4, 95% CI=0.1–1.4). In the cohort study of Finnish workers exposed to xylene, toluene, and styrene, Anttila and colleagues (1998) found increased risks of pancreatic cancer (SIR=1.26, 95% CI=0.41–2.93). A case-control study of pancreatic cancer by Ji and colleagues (1999) conducted in Shanghai showed increased risks in various occupational groups, especially among male painters (OR=5.2, 95% CI=1.1–25.0).

The comprehensive cohort study of methylene chloride-exposed workers at Kodak has been followed for a number of years (Hearne et al., 1987, 1990). In the initial publication, an excess of pancreatic cancer was observed (SMR=3.1); however, in the second study, after 4 additional years of followup, there was no increase in pancreatic cancer mortality. In a study examining the same cohort and another cohort of Kodak workers, Hearne and Pifer (1999) showed an increased risk of pancreatic cancer associated with career exposure to methylene chloride of over 800 ppm-years on the basis of three cases (SMR=2.34, compared with internal controls). Among workers who were employed in 1964–1970 in the roll coating division, the highest risk was found in the lowest cumulative-exposure category of less than 400 ppm (SMR=2.58, compared with internal controls). Other cohorts of methylene chloride-exposed workers had very few cases of pancreatic cancer and reported no increased risk (Gibbs et al., 1996; Lanes et al., 1990, 1993; Tomenson et al., 1997).

The study of workers in five chemical plants found no increased risk of pancreatic cancer associated with exposure to phenol (SMR=0.6, 95% CI=0.4–1.1) (Dosemeci et al., 1991).

Several studies reported associations between pancreatic cancer and unspecified mixtures of organic solvents (Table 6.11). The studies showing positive associations included male leather workers in Italy (SMR=1.46, 95% CI=0.39–3.73) (Costantini et al., 1989), aircraft-manufacturing workers (SMR=1.19, 95% CI=0.83–1.67) (Garabrant et al., 1988), and the

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

case-control study in Finland of high exposure to solvents (OR=2.01, 95% CI=0.98–4.10) (Kauppinen et al., 1995).

Summary and Conclusion

For exposure to trichloroethylene, tetrachloroethylene and dry-cleaning solvents, xylene, toluene, benzene, methylene chloride, phenol, and solvent mixtures and the risk of pancreatic cancer, the evidence was limited by mixed results, the lack of exposure-response relationships, and imprecise estimates of risk. Table 6.11 identifies the studies reviewed by the committee in making its conclusion regarding association. All of the study populations include both men and women unless stated otherwise.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to solvents under review and pancreatic cancer.

TABLE 6.11 Selected Epidemiologic Studies—Pancreatic Cancer and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Studies—Incidence

Hansen et al., 2001

Biologically monitored Danish workers

3

1.0 (0.2–3.0)

Blair et al., 1998

Aircraft maintenance workers in Utah

 

 

No trichloroethylene exposure

6

0.7 (0.2–2.3)

 

<5 unit-years

6

0.7 (0.2–2.1)

 

5–25 unit-years

2

0.4 (0.1–1.8)

 

>25 unit-years

51

0.7 (0.2–2.4)

Anttila et al., 1995

Biologically monitored Finnish workers

 

 

Entire period since first measurement:

11

1.61 (0.81–2.88)

 

0–9 years

1

0.56 (0.01–3.10)

 

10–19 years

8

2.30 (0.99–4.52)

 

20+ years

2

1.31 (0.16–4.74)

 

Mean personal U-TCA level:

 

 

<100 µmol/L

6

1.61 (0.59–3.50)

 

100+ µmol/L

3

1.31 (0.27–3.82)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California, potential routine exposure

7

0.41 (0.17–0.85)

Blair et al., 1998

Aircraft-maintenance workers in Utah, employed >1 year

33

1.2 (0.6–2.3)

Morgan et al., 1998

Aerospace workers in Arizona

 

Trichloroethylene-exposed subcohort

11

0.76 (0.38–1.37)

 

High trichloroethylene-exposure

6

0.66 (0.24–1.43)

Tetrachloroethylene and Dry-cleaning Solvents

Cohort Study—Incidence

Anttila et al., 1995

Biologically monitored Finnish workers

3

3.08 (0.63–8.99)

Cohort Study—Mortality

Ruder et al., 2001

Dry-cleaning labor-union workers

18

1.53 (0.91–2.42)

 

Tetrachloroethylene-only

3

0.80 (0.17–2.35)

 

Tetrachloroethylene-plus other solvents

15

1.89 (1.06–3.11)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Xylene, Toluene, Benzene

Cohort Study—Incidence

Anttila et al., 1998

Finnish workers biologically monitored for exposure to aromatic hydrocarbons (styrene, toluene, xylene)

5

1.26 (0.41–2.93)

Case-Control Studies

Ji et al., 1999

Residents of Shanghai, China

 

 

Chemical and rubber workers (female)

5

1.4 (0.4–4.7)

 

Rubber workers (female)

5

1.7 (0.5–5.8)

 

Printers (male)

4

5.2 (1.1–25.0)

Gérin et al., 1998

Male residents of Montreal, Canada

 

 

Xylene, medium/high exposure

4

1.1 (0.4–3.3)

 

Toluene, medium/high exposure

3

0.6 (0.2–2.2)

 

Benzene, medium/high exposure

3

0.4 (0.1–1.4)

Methylene Chloride

Cohort Studies—Mortality

Hearne and Pifer, 1999

Male Kodak workers in New York state, employed >1 year

 

Methylene chloride cohort

5

0.92 (0.30–2.14)

 

Internal comparison, ≥800 ppm-years

3

2.34

 

Roll-coating division (New York state control)

8

1.51 (0.65–2.98)

 

Roll-coating division (Kodak Rochester control)

8

1.55 (0.67–3.06)

Tomenson et al., 1997

Male cellulose triacetate film workers, ever employed

3

0.68 (0.14–1.99)

Gibbs et al., 1996

Cellulose-fiber production workers

 

 

High exposure, males

1

0.35 (0.01–1.92)

 

High exposure, females

0

0

 

Low exposure, males

2

0.89 (0.11–3.22)

 

Low exposure, females

1

0.58 (0.01–3.23)

Lanes et al., 1993

Cellulose-fiber production plant workers, employed >3 months

2

0.83 (0.10–2.99)

Phenol

Cohort Study—Mortality

Dosemeci et al., 1991

Male workers in five US chemical plants

 

Phenol, any exposure

14

0.6 (0.4–1.1)

Unspecified Mixtures of Organic Solvents

Cohort Study—Incident

Anttila et al., 1995

Finnish workers biologically monitored for exposure to halogenated hydrocarbons (trichloroethylene, tetrachloroethylene, 1,1,1-trichloroethane)

12

1.56 (0.81–2.72)

Cohort Studies—Mortality

Fu et al., 1996

Shoemakers in England and Florence

 

 

English cohort

25

0.70 (0.45–1.04)

 

Florence cohort

2

0.54 (0.07–1.95)

Acquavella et al., 1993

Metal-components manufacturing workers, ever exposed

1

2.9 (0.1–16.0)

Costantini et al., 1989

Male leather workers in Tuscany, Italy, employed >6 months

4

1.46 (0.39–3.73)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Garabrant et al., 1988

Aircraft-manufacturing workers in California, employed >4 years

34

1.19 (0.83–1.67)

Pippard and Acheson, 1985

Male boot and shoe makers in three English towns

 

Rushden

21

0.96 (0.59–1.46)

 

Street

2

0.40 (0.05–1.46)

 

Stafford

6

0.96 (0.35–2.11)

McMichael et al., 1976

Male rubber workers in Ohio and Wisconsin

 

Age 40–64

17

0.95 (0.55–1.52)a

 

Age 65–84

40

1.08 (0.77–1.47)a

 

Age 40–84

57

1.03 (0.78–1.33)a

Case-Control Study

Kauppinen et al., 1995

Residents of Finland

 

Solvents (all)

20

1.22 (0.73–2.07)

 

Solvents (high)

14

2.01 (0.98–4.10)

NOTE: U-TCA=urinary metabolite of trichloroethylene.

a95% CI calculated by the committee with standard methods from the observed and expected numbers presented in the original study.

HEPATOBILIARY CANCERS

Description of Case-Control Studies

Two population-based (Hardell et al., 1984; Heinemann et al., 2000) and two hospital-based (Hernberg et al., 1988; Stemhagen et al., 1983) case-control studies examined risk of liver cancer associated with occupational exposure to solvents (Table 6.12). Self-administered questionnaires or interviews were used to obtain occupational history information in each study. The study by Stemhagen and colleagues (1983) used job titles as surrogates of exposure. In two of the other studies, exposures were inferred by industrial hygienists (Heinemann et al., 2000; Hernberg et al., 1988). In the two hospital-based studies, there were no adjustments for potentially confounding variables.

Epidemiologic Studies of Exposure to Organic Solvents and Hepatobiliary Cancers

Anttila and colleagues (1995) found an increased risk (SIR=2.27, 95% CI=0.74–5.29) of liver cancer in the cohort of workers biologically monitored for a metabolite of trichloroethylene. Likewise, Hansen and colleagues (2001) reported an increased risk of cancer of the liver and biliary passages in men in a Danish cohort of biologically monitored workers (SIR=2.6, 95% CI=0.8–6.0).

In a large cohort study of aircraft-manufacturing workers in California, Boice and colleagues (1999) found no association between liver cancer and exposure to trichloroethylene (SMR=0.54, 95% CI=0.15–1.38). Another cohort of aircraft manufacturers (Morgan et al., 1998) showed no increased mortality in trichloroethylene-exposed workers, and there was no evidence of a trend in mortality with cumulative exposure.

In the cohort of workers at Hill Air Force Base in Utah, Blair and colleagues (1998) did not find excess risk of liver cancer mortality or incidence, nor was there any apparent increase with increasing cumulative exposure to trichloroethylene.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.12 Description of Case-Control Studies of Liver Cancer and Exposure to Organic Solvents

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Stemhagen et al., 1983

Cases identified through diagnosis in New Jersey hospitals in 1975–1980 or from death certificates in 1975–1979, all with histologic confirmation; controls selected from hospital records and death certificates and matched for age, race, sex, county of residence

265

530

Laundering, cleaning, other garment service work

In-person interview (direct or proxy) to assess occupational history (job titles)

OR

None

Response rates: 89.5% of cases, 77% of controls

Hardell et al., 1984

Deceased male cases, age 25–80 years at diagnosis, reported to the Swedish Cancer Registry and diagnosed in 1974–1981, with histologic confirmation; controls selected from the National Population Register, matched on sex, age, year of death, municipality

98

200

Organic solvents

Mailed questionnaires to next of kin, assessing work history (job titles) and occupational or leisure-time exposure to specific chemicals (self-reports)

Mantel-Haenszel rate ratio, Miettinen 95% CI

Age

Hernberg et al., 1988

Deceased cases, reported to the Finnish Cancer Register in 1976–1978 and 1981; controls selected from deceased stomach-cancer cases and coronary-infarction patients reported in 1977

344

476 stomach

385 coronary infarction

Solvents

Questionnaire mailed to next of kin to elicit list of occupations and employers; hygienist assigned exposure classification based on occupation

Likelihood-based ORs

None

Response rates: 71.7% of cases, 72.8% of stomach-cancer controls, 69.0% of infarction controls

Heinemann et al., 2000

Female cases, under 65 years old, identified at 64 clinics in six European countries in 1990–1994 (prevalent cases) and 1994–1996 (incident cases); hospital controls selected from respective clinics and matched on age; population controls selected from citizen registers

317

1,779

Dry-cleaning work Solvents

In-person interview (direct or proxy) with questionnaire assessing occupational history (industry titles) and specific agent exposures (self-reports)

Logistic regression

Age, center, smoking, alcohol, oral contraceptive use, hepatitis infection

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Ritz (1999) found associations between exposure to trichloroethylene and liver cancer among male workers in the uranium-processing industry (SMR=1.66, 95% CI=0.71–3.26). The association is based on a small number of cases and may be confounded by other potential exposures involved in uranium processing. Axelson and colleagues (1994) also found an increased incidence in male Swedish workers exposed to trichloroethylene (SIR=1.41, 95% CI =0.38–3.60). A cohort study of transformer-assembly workers (Greenland et al., 1994) showed no positive associations between occupational exposure to trichloroethylene and cancers of the liver, gallbladder, and biliary tract combined (OR=0.54, 95% CI=0.11–2.63).

In the cohort of aircraft-manufacturing workers (Boice et al., 1999), an association between exposure to tetrachloroethylene and liver cancer was observed (SMR=2.05, 95% CI=0.83–4.23) but no increase was observed with increasing duration of exposure. Bond and colleagues (1990) found an association between mortality and liver cancer in a cohort of chemical workers (OR=1.8, 95% CI=0.8–4.3).

A cohort study of US dry cleaners showed no association to liver cancer (SMR=0.7, 95% CI=0.2–1.7) (Blair et al., 1990). Ruder and colleagues (2001) detected only one case of liver cancer among the subcohort of workers exposed to tetrachloroethylene and other solvents (SMR=0.16, 95% CI=0.00–1.32).

In a case-control study of women in the Multicentre International Liver Tumour Study, three women with hepatocellular cancer reported working as dry cleaners (OR=0.65; 95% CI=0.12–3.44) (Heinemann et al., 2000). Another study of occupational risk factors for liver cancer found an association among men employed in laundering, cleaning, and other garment services (RR=2.5, 95% CI=1.02–6.14) (Stemhagen et al., 1983). Further investigation by the authors showed that the cases were concentrated among people who processed clothes and potentially had exposure to other chemicals.

Friedlander and colleagues (1978) established a cohort of workers in one department at Kodak where methylene chloride was the primary solvent exposure for more than 30 years. In the most recent followup (Hearne and Pifer, 1999), one death from liver cancer was observed (SMR=0.42, 95% CI=0.01–2.36).

A cohort study of workers producing cellulose triacetate fibers at a Hoechst Celanese plant in South Carolina showed an excess risk of biliary and liver cancers (SMR=2.98, 95% CI =0.81–7.63) (Lanes et al., 1993). In another Hoechst Celanese facility manufacturing cellulose triacetate, Gibbs and colleagues (1996) found one death from liver cancer (SMR=0.81, 95% CI =0.02–4.49).

No cases of liver cancer were found in a cohort of male workers producing cellulose triacetate film base at a plant in the UK (Tomenson et al., 1997). A nested case-control study of liver and biliary tract cancer cases identified among male hourly employees of the Dow Chemical Company did not show an excess risk in workers exposed to methylene chloride (RR=0.8, 95% CI=0.2–3.6) (Bond et al., 1990).

Yin and colleagues (1996a,b) followed a cohort of factory workers with known exposure to benzene. Increased relative risks of cancers of the liver and gallbladder combined were reported for men (RR=1.3, 95% CI=0.9–1.9) but not for women (RR=0.4, 95% CI=0.2–1.3). This cohort was further examined by Hayes and colleagues (1996), who evaluated the relative risks according to cumulative exposure to benzene. There was some suggestion that the relative risks for liver and gallbladder cancer increased with increasing cumulative exposure, although chance could not be ruled out confidently (p value for linear trend=0.16).

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

A nested case-control study by Greenland and colleagues (1994) of transformer-assembly workers exposed to benzene showed a positive association between occupational exposure to benzene and cancers of the liver, gallbladder, and biliary tract combined (OR=2.76, 95% CI=0.68–11.2).

A cohort of toluene-exposed German rotogravure workers studied by Wiebelt and Becker (1999) experienced a higher risk of liver cancer mortality than the population of West Germany (SMR=1.98, 95% CI=0.34–7.16). Dosemeci and colleagues (1991) conducted a study of US industrial workers that showed no increase in liver cancer mortality among those occupationally exposed to phenol (SMR=1.0, 95% CI=0.4–1.9). The association with phenol did not increase markedly from low to medium to high exposure.

Three population-based case-control studies (Hardell et al., 1984; Heinemann et al., 2000; Hernberg et al., 1988) examined the association of liver cancer in relation to the broad category of solvents or mixed solvents. Heinemann and colleagues (2000) found no associations (OR=1.05, 95% CI=0.52–2.09). Hernberg and colleagues (1988) did not find increased risks among males (OR=0.6, 90% CI 0.3–1.3), but increased risks were found among women (OR=3.4, 90% CI 1.3–8.6). Hardell and colleagues (1984) also found an association (OR=1.8, 95% CI=0.99–3.4).

In the cohort studies, no associations were found in aircraft-manufacturing workers (SMR =0.92) (Boice et al., 1999), but the three cohort studies of painters all showed excess risks of liver cancer: Steenland and Palu (1999) found an SMR of 1.25 (95% CI=1.03–1.50), Matanoski and colleagues (1986) an SMR of 1.47 (95% CI=0.98–2.13), and Morgan (1981) an SMR of 1.93.

Summary and Conclusion

Although most studies examined the risk for liver cancer broadly, others combined liver cancer with gallbladder and biliary tract cancers. For exposure to trichloroethylene, tetrachloroethylene and dry-cleaning solvents, toluene, and phenol, most studies did not find an increased risk. Although one study on methylene chloride found a positive association, the small number of cases of liver cancer and the lack of corroborating studies were limitations of the literature. The studies on exposure to benzene were also limited with only three studies of benzene-exposed workers (two of which were on the same cohort), and only weak evidence of excess risks. For exposure to unspecified mixtures of organic solvents, although some of the estimates of risk were positive, others were not, and the role of confounding by other exposures or risk factors was a limitation. Table 6.13 identifies the key studies and the relevant data points reviewed by the committee in drawing its conclusion. Unless indicated in the table, the study populations include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to solvents under review and hepatobiliary cancers.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.13 Selected Epidemiologic Studies—Hepatobiliary Cancers and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Studies—Incidence

Hansen et al., 2001

Biologically monitored Danish workers

 

 

Males

5

2.6 (0.8–6.0)

 

Females

0

Blair et al., 1998

Aircraft-maintenance workers in Utah

 

 

Males:

 

 

No exposure

1

0.8 (0.1–12.0)

 

<5 unit-years

2

1.2 (0.1–13.8)

 

5–25 unit-years

1

1.0 (0.1–16.0)

 

>25 unit-years

3

2.6 (0.3–25.0)

Anttila et al., 1995

Biologically monitored workers in Finland

 

 

Entire period since first measurement:

5

2.27 (0.74–5.29)

 

0–9 years

0

—(0.0–6.59)

 

10–19 years

2

1.74 (0.21–6.29)

 

≥20 years

3

6.07 (1.25–17.7)

 

Mean personal U-TCA level

 

 

<100 µmol/L

2

1.64 (0.20–5.92)

 

100+ µmol/L

2

2.74 (0.33–9.88)

Axelson et al., 1994

Biologically monitored Swedish workers

4

1.41 (0.38–3.60)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

All exposed factory workers

4

0.54 (0.15–1.38)

 

Duration of potential exposure (routine or intermittent)

 

 

<1 year

4

0.53 (0.18–1.60)

 

1–4 years

3

0.52 (0.15–1.79)

 

≥5 years

6

0.94 (0.36–2.46)

Ritz, 1999

White male US uranium-processing plant workers

 

 

Trichloroethylene, cutting fluids, or kerosene

8

1.66 (0.71–3.26)

 

Trichloroethylene—light exposure

 

 

>2 years, no latency

3

0.93 (0.19–4.53)

 

>2 years, 15-year latency

3

1.16 (0.24–5.60)

 

>5 years, no latency

3

1.90 (0.35–10.3)

 

>5 years, 15-year latency

3

2.86 (0.48–17.3)

 

Trichloroethylene—moderate exposure

 

 

>2 years, no latency

1

4.97 (0.48–51.1)

 

>2 years, 15-year latency

1

5.53 (0.54–56.9)

 

>5 years, no latency

1

8.82 (0.79–98.6)

 

>5 years, 15-year latency

1

12.1 (1.03–144)

Blair et al., 1998

Aircraft-maintenance workers in Utah

 

 

Males

 

 

No exposure

3

0.5 (0.1–2.4)

 

<5 unit-years

6

1.1 (0.3–4.1)

 

5–25 unit-years

3

0.9 (0.2–4.3)

 

>25 unit-years

3

0.7 (0.2–3.2)

 

Females:

 

 

No exposure

3

4.2 (0.7–25.0)

 

<5 unit-years

1

1.6 (0.2–18.2)

 

5–25 unit-years

0

 

>25 unit-years

2

2.3 (0.3–16.7)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Morgan et al., 1998

Aerospace workers in Arizona

 

 

Trichloroethylene-exposed subcohort:

6

0.98 (0.36–2.13)

 

Low cumulative exposure

3

1.32 (0.27–3.85)

 

High cumulative exposure

3

0.78 (0.16–2.28)

 

Peak and cumulative exposure:a

 

 

Peak: medium and high vs low and no exposure

3

0.98 (0.29–3.35)

 

Cumulative (low)

3

2.12 (0.59–7.66)

 

Cumulative (high)

3

1.19 (0.34–4.16)

Greenland et al., 1994

White male transformer-assembly workers, ever exposed

NA

0.54 (0.11–2.63)

Tetrachloroethylene and Dry-cleaning Solvents

Cohort Studies—Mortality

Ruder et al., 2001

US dry-cleaning workers in four labor unions

1

0.16 (0.00–1.32)

Boice et al., 1999

Aircraft- manufacturing workers

 

 

All exposed factory workers

7

2.05 (0.83–4.23)

 

Duration of potential exposure (routine or intermittent)

 

 

<1 year

3

1.38 (0.40–4.69)

 

1–4 years

4

0.39 (0.39–3.47)

 

≥5 years

5

1.29 (0.46–3.65)

Blair et al., 1990

Dry-cleaning union members in Missouri

5

0.7 (0.2–1.7)

Bond et al., 1990

Male chemical-company workers in Michigan, ever exposed

6

1.8 (0.8–4.3)

Case-Control Studies

Heinemann et al., 2000

Females in the Multicentre International Liver Tumour Study

 

 

Dry-cleaning, ever employed

3

0.65 (0.12–3.44)

Stemhagen et al., 1983

Laundering, cleaning, other garment-services workers in New Jersey, employed >6 months

10

2.50 (1.02–6.14)

Methylene Chloride

Cohort Studies—Mortality

Hearne and Pifer, 1999

Male cellulose triacetate photographic-film base workers in Kodak Park, employed >1 year

1

0.42 (0.01–2.36)

Tomenson et al., 1997

Male cellulose triacetate-fiber film base workers in the UK

 

 

Never exposed

0

 

All exposed

0

Gibbs et al., 1996

Male cellulose triacetate-fiber production workers in Maryland

 

 

No exposure

0

 

Low exposure

1

0.75 (0.029–4.20)

 

High exposure

1

0.81 (0.020–4.49)

Lanes et al., 1993

Cellulose triacetate-fiber production workers in South Carolina, employed >3 months

4

2.98 (0.81–7.63)

Bond et al., 1990

Male chemical-company workers in Michigan, ever exposed

2

0.8 (0.2–3.6)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Benzene

Cohort Studies—Mortality

Yin et al., 1996a

Chinese factory workers, ever exposed

 

 

Total cohort

109

1.2 (0.8–1.6)

 

Men

101

1.3 (0.9–1.9)

 

Women

8

0.4 (0.2–1.3)

Hayes et al., 1996

Chinese factory workers

 

 

<10 ppm-years

12

1.1

 

10–39 ppm-years

12

0.8

 

40–99 ppm-years

9

0.6

 

100–400 ppm-years

44

1.6

 

400+ ppm-years

28

1.2

 

 

 

p-trend=0.16

Greenland et al., 1994

White male transformer-assembly workers, ever exposed

NA

2.76 (0.68–11.2)

Other Specific Organic Solvents

Cohort Studies—Mortality

Wiebelt and Becker, 1999

Male German rotogravure printing-plant workers, employed >1 year—toluene

3

1.98 (0.34–7.16)

Dosemeci et al., 1991

US white male industrial workers—phenol

 

 

Any exposure

8

1.0 (0.4–1.9)

 

Level of cumulative exposure:

 

 

None

4

1.2

 

Low exposure

1

0.3

 

Medium exposure

6

1.6

 

High exposure

1

1.4

Unspecified Mixtures of Organic Solvents

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers

 

 

All exposed factory workers

17

0.92 (0.54–1.47)

 

Duration of potential exposure (routine or intermittent)

 

 

<1 year

10

1.35 (0.63–2.87)

 

1–4 years

14

0.75 (0.38–1.47)

 

≥5 years

31

0.97 (0.54–1.72)

Steenland and Palu, 1999

Members of US painters’ unions

 

 

Total cohort

119

1.25 (1.03–1.50)

 

20 years since first union membership

90

1.17 (0.95–1.44)

Greenland et al., 1994

White male transformer-assembly workers, ever exposed to solvents

NA

0.69 (0.18–2.60)

Matanoski et al., 1986

US painters and allied tradesmen

28

1.47 (0.98–2.13)

Morgan et al., 1981

Male US paint or varnish manufacturing workers, employed >1 year

6

1.93

Case-Control Studies

Heinemann et al., 2000

Women in Multicentre International Liver Tumor Study, ever exposed to solvents

18

1.05 (0.52–2.09)

Hernberg et al., 1988

Finnish cases and deceased controls

 

 

Solvent exposure, 10-year latency

 

 

Males

7

0.6 (0.3–1.3)b

 

Females

7

3.4 (1.3–8.6)b

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Hardell et al., 1984

Male residents of Sweden

 

 

High-grade exposure to organic solvents

 

 

Primary liver cancer

22

1.8 (0.99–3.4)

 

Hepatocellular carcinoma

20

2.1 (1.1–4.0)

NOTE: NA=not available; U-TCA=urinary metabolite of trichloroethylene.

aInternal cohort analyses for peak and cumulative exposure to trichloroethylene classifications used Cox proportional-hazards models.

b90% CI.

LUNG CANCER

Description of Case-Control Studies

Three population-based case-control studies (Table 6.14) reported the risk of lung cancer associated with job title and self-reported exposure to specific chemicals (Brownson et al., 1993; Pohlabeln et al., 2000) or to tetrachloroethylene in drinking water (Paulu et al., 1999). Smoking is a known risk factor for lung cancer and was accounted for in each of the studies (see Chapter 2 and Appendix E for more information on smoking).

Epidemiologic Studies of Exposure to Organic Solvents and Lung Cancer

No association was found between incidence of lung cancer and concentrations of the biologic marker of exposure to trichloroethylene (SIR=0.92, 95% CI=0.59–1.35) (Anttila et al., 1995). A cohort of aircraft-manufacturing workers with potential exposure to trichloroethylene was followed for nearly 30 years, but no association between lung cancer and trichloroethylene was reported (SMR=0.76, 95% CI=0.60–0.95) (Boice et al., 1999). The authors noted that most workers were exposed to a variety of substances routinely or intermittently. A mortality study of civilian aircraft-maintenance workers at Hill Air Force Base in Utah that included an extensive assessment of exposure to trichloroethylene did not show an increase in lung cancer mortality or incidence in men (over 25 unit-years: SMR=1.1, 95% CI=0.7–1.8) or women (over 25 unit-years: SMR=0.4, 95% CI=0.1–1.8) (Blair et al., 1998). The cohort study of aircraft manufacturers in Arizona (Morgan et al., 1998) did not show increased mortality from lung cancer in the trichloroethylene-exposed subcohort (SMR=1.10, 95% CI=0.89–1.34).

In addition to those studies, no associations were found for rubber-industry workers exposed to trichloroethylene (OR=0.64) (Wilcosky et al., 1984) or to transformer-assembly workers (OR=1.01, 95% CI=0.69–1.47) (Greenland et al., 1994).

A cohort study of dry-cleaning workers exposed to tetrachloroethylene and other solvents showed increased mortality from lung cancer (SMR=1.46, 95% CI=1.07–1.95) (Ruder et al., 2001). In the subcohort of workers exposed only to tetrachloroethylene, the relative risks were not as great (SMR=1.17, 95% CI=0.71–1.83). Blair and colleagues (1990) reported a comparable relative risk of lung cancer in another cohort of dry cleaners (SMR=1.3, 95% CI=0.9–1.7).

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.14 Description of Case-Control Studies of Lung Cancer and Exposure to Organic Solvents

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Brownson et al., 1993

White, female cases, age 30–84 years, identified through the Missouri Cancer Registry as diagnosed in 1986–1991, with 77% histologic confirmation; controls younger than 65 years selected from state driver’s license files; controls older than 65 years were selected from the records of HCFA

429

1,021

Dry-cleaning work

Printing-industry work

Trained interviewers conducted telephone and in-person interviews (direct or proxy) to assess specific occupational (job titles) and exposure history (self-reported)

Multiple logistic regression

Age, history of lung disease, active smoking

Response rates: 66% of cases, 67% of controls

Paulu et al., 1999

Cases reported to the Massachusetts Cancer Registry, diagnosed in 1983–1986 among residents of five upper Cape Cod towns; living controls were selected from the records of HCFA and through RDD; deceased controls identified by the state Department of Vital Statistics and Research files

243

1,206

Tetrachloroethylene

Exposure dose estimated in areas of contaminated drinking water, accounting for location and years of residence, water flow, pipe characteristics

Multiple logistic regression

Age at diagnosis, vital status, sex, occupational exposure to solvents; specific cancer risk factors controlled in respective analyses, such as smoking

Response rates: 79% of cases, 76% of HCFA controls, 74% of RDD controls, 79% of next of kin of deceased controls

Pohlabeln et al., 2000

Nonsmoking cases identified in 12 study centers from seven European countries between 1988–1994 (96.5% were histologically confirmed); community-based controls were selected in six centers, hospital-based controls in five centers, and both community and hospital-based controls in one center

650

1,542

Laundry and dry cleaners

Interviewers used a common questionnaire to record lifelong occupational histories which were coded by job titles

Unconditional logistic regression

Sex, occupation, age, center, occasional smoking (ever smoked occasionally, but fewer than 400 cigarettes), residence, diet, and environmental tobacco smoke

Response rates: across centers the rate varied between 55% and 95%; in three centers, the response rate among control subjects was lower than 50%

NOTE: HCFA=Health Care Financing Administration; RDD=random-digit dialing.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

An increased risk of lung cancer was found in workers exposed to tetrachloroethylene (SIR=1.92, 95% CI=0.62–4.48) (Anttila et al., 1995). Among the cohort of aircraft manufacturers, no association was found between exposure to tetrachloroethylene and lung cancer (SMR=1.08, 95% CI=0.79–1.44) (Boice et al., 1999). The authors indicated that concentrations of tetrachloroethylene in the air samples were relatively low and that exposures were well below permissible concentrations. Wilcosky and colleagues (1984), in a study of rubber-industry workers, reported no increased risk of lung cancer with exposure to tetrachloroethylene.

The risk of developing lung cancer in relation to exposure to tetrachloroethylene was evaluated in the two case-control studies, and these were adjusted for smoking. A risk associated with self-reported exposure to tetrachloroethylene was found among lifetime nonsmokers (OR=2.1, 95% CI=1.2–3.7) (Brownson et al., 1993). A case-control study of residents of upper Cape Cod (Paulu et al., 1999) showed excess risks of lung cancer with increasing level of estimated exposure to tetrachloroethylene in drinking water. A third case-control study of female, nonsmoking laundry and dry cleaners also found an increased risk of lung cancer (OR=1.83, 95% CI=0.98–3.40) (Pohlabeln et al., 2000).

Several cohort studies of workers exposed to methylene chloride provided little support for an association between exposure and lung cancer (Gibbs et al., 1996; Hearne and Pifer, 1999; Hearne and Friedlander, 1981; Hearne et al., 1987, 1990; Lanes et al., 1990, 1993; Tomenson et al., 1997). Those large cohort studies followed methylene chloride-exposed workers for many years, some with repeated followup, and examined the association between exposure and cancer mortality, but they reveal no excess of lung cancer associated with exposure.

Yin and colleagues (1996a,b) reported increased relative risks of cancers of the trachea, bronchi, and lung combined in benzene-exposed males (RR=1.5, 95% CI=1.0–2.2) and in the entire cohort of exposed men and women (RR=1.4, 95% CI=1.0–2.0); no association was found in benzene-exposed women (RR=1.0, 95% CI=0.4–2.9). Hayes and colleagues (1996) assessed the cumulative exposure to benzene in the overall cohort and found that the relative risks of tracheal, bronchial, and lung cancers combined increased with increasing exposure (p trend=0.01).

Two nested case-control studies, one of transformer-assembly workers (Greenland et al., 1994) and another of rubber-industry workers (Wilcosky et al., 1984), did not show any associations between occupational exposure to benzene and lung cancer. No data were presented on increasing levels of exposure.

Three studies provided evidence on the association between toluene and lung cancer. A cohort of toluene-exposed German rotogravure workers showed increased lung cancer mortality when compared with the mortality in West Germany (SMR=1.23, 95% CI=0.81–1.92) (Wiebelt and Becker, 1999). Job-specific subcohorts with different levels of exposure to toluene demonstrated a range of lung cancer risk, from no risk in printing-cylinder preparation occupations to an SMR of 1.77 (95% CI=0.77–4.39) in finishing workers who had the lowest level of exposure over the entire observation period.

A Swedish cohort of rotogravure printers showed similar increases in respiratory tract cancer mortality from exposure to toluene (SMR=1.76, 95% CI=1.03–2.91) (Svensson et al., 1990); however, no gradient was found with duration of exposure.

Factory workers who were considered to have high exposure in a plant manufacturing chlorinated toluenes (benzyl chloride, benzal chloride, benzotrichloride, and benzoyl chloride) experienced increased lung cancer mortality (SMR=3.31, 95% CI=1.59–6.09) (Sorahan and

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Cathcart, 1989), and there was evidence that risk increased with increasing exposure. In a nested case-control study of 26 cases of lung cancer from the cohort of chlorinated toluene production workers, RR of lung cancer associated with benzotrichloride and “other chlorinated toluenes” was 1.36 (95% CI=0.43–24) and 1.12 (95% CI=0.30–4.22), respectively, per 10 years of exposed employment. In a study of rubber workers, Wilcosky and colleagues (1984) did not find an association (OR=0.55).

There were three studies of the association between exposure to phenol and lung cancer. A cohort study with many exposed cases showed no association (Dosemeci et al., 1991), and no association was reported for workers in the rubber industry (Wilcosky et al., 1984). Kauppinen and colleagues (1993) reported a four-fold excess risk (SMR=4.04, 95% CI=1.83–8.89) after adjusting for smoking among those exposed for at least 1 month.

Several other solvents used in the Gulf War—including naphtha, ethanol, xylenes, isopropanol, ethyl acetate, and acetone (Wilcosky et al., 1984)—were investigated in relation to lung cancer, and no associations with these solvents were found. Anttila and colleagues (1995) found an increase in lung and bronchial cancer risk with exposure to 1,1,1-trichloroethane (SIR=1.31, 95% CI=0.16–4.71).

Several studies, including some described previously, examined the association of lung cancer with unspecified mixtures of solvents. Most studies did not identify “solvents” or “organic solvents” as the exposure being evaluated but instead defined exposure by various occupational titles or groups, such as painters (Engholm and Englund, 1982; Englund 1980; Matanoski et al., 1986; Morgan et al., 1981; Steenland and Palu, 1999; Stockwell and Matanoski, 1985), printers (Malker and Gemne, 1987), workers in transformer assembly (Greenland et al., 1994), ethanol and isopropanol production (Teta et al., 1992), isopropanol and methyl ethyl ketone production (Alderson and Rattan, 1980), shoe manufacturing (Walker et al., 1993), chemical manufacturing (Waxweiler et al., 1981), aircraft manufacturing (Boice et al., 1999), or metal-component manufacturing (Acquavella et al., 1993). As shown in Table 6.15, many of the studies showed positive associations between occupational exposure and lung cancer.

Generally, workers in the studies were exposed to solvents and other chemical agents not reviewed by this committee. More important, although smoking is associated with lung cancer, no adjustments for it were made in most cohort studies. (That is not unusual; obtaining smoking histories, especially in mortality studies, is difficult.) Furthermore, one study that specifically examined lung cancer risk among nonsmokers in the printing industry found no association with occupational exposure (Brownson et al., 1993). Unlike farm workers, who have some of the lowest smoking rates in the United States, painters, truck drivers, construction workers, carpenters, and auto mechanics have some of the highest rates. For example, in a National Center for Health Statistics survey of men 20–64 years old in 1978–1980, 55.1% of painters surveyed reported that they smoked (US Surgeon General, 1985).

To understand the impact of smoking on risk of lung cancer, the committee examined the risk estimates for bladder cancer and cardiovascular disease—diseases for which smoking is also a known risk factor. Almost all the studies reported rates of bladder cancer and cardiovascular disease similar to those of lung cancer. Almost all the studies also stated that asbestos was present in the workplace and probably contributed to the slightly increased observed risks of lung cancer. As a result, the committee determined that exposure to solvents alone was an unlikely explanation for the increased risks of lung cancer and that confounding by smoking was possibly biasing the results.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Summary and Conclusion

Although there are different types of respiratory cancers, most studies assessed exposure in relation to lung cancer. Exposures were typically defined by occupations in which the solvents were known to be present; and confounding factors, especially smoking, were not consistently controlled for in the analyses. In reviewing the literature, the committee found that there was no evidence from any of the studies of a positive association between exposure to trichloroethylene, methylene chloride, benzene, phenol, and other specific solvents and risk of lung cancer.

For exposure to tetrachloroethylene and dry-cleaning solvents, some committee members believed that the overall evidence was limited by the possibility of confounding, and that most findings were based on small numbers of cases exposed by different routes. The cohort studies did not control for other occupational exposures or smoking, an important potential confounder for lung cancer. However, other committee members believed that the consistently positive findings and evidence of a dose-response relationship in the case-control study by Paulu and colleagues were supportive of a conclusion of limited/suggestive evidence. Both case-control studies adjusted for smoking and still found relatively high relative risks of lung cancer. As a result, the committee decided not to state a formal consensus conclusion. Additional studies that control for smoking and address other concerns related to misclassification of exposure are needed before a more definitive conclusion as to exposure to tetrachloroethylene and dry-cleaning solvents and the risk of lung cancer can be reached.

Although several cohort studies reported increased risk of lung cancer associated with exposure to toluene, estimates were weak. Workers were probably exposed to other compounds that were not controlled for in the analyses. In addition, information on smoking was not available. A unique relationship between solvent exposure and lung cancer was not found, and the committee concluded that the evidence was inadequate/insufficient to support an association between exposure to unspecified mixtures of organic solvents and lung cancer.

Several studies on specific organic solvents and solvent mixtures found positive associations between exposure and the risk of lung cancer. However, most studies on specific solvents were too small and inconsistent in their findings to support conclusions. Some studies showed positive associations, but they were limited by lack of information on smoking habits among cohort members.

With respect to exposure to solvent mixtures, many studies reported positive findings; most, however, were based on occupational titles or industries and lacked specific analyses of “solvents” or “organic solvents.” Although their results suggested a possible relationship, the lack of smoking data, the lack of exposure specificity, and the potential for confounding by other occupational exposures (such as to asbestos) limited their utility. Future research with sufficient power would help to clarify whether an association between exposure to solvent mixtures or the interactions of various solvents and lung cancer exists, as is indicated in some of the studies reviewed by the committee. Table 6.15 identifies the key studies and the relevant data points evaluated by the committee in drawing its conclusion. Unless indicated in the tables, the study populations include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to solvents under review, other than tetrachloroethylene and dry-cleaning solvents, and lung cancer.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.15 Selected Epidemiologic Studies—Lung Cancer and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Studies—Incidence

Blair et al., 1998

Aircraft-maintenance workers in Utah

 

 

Males

 

 

No exposure

22

1.0 (0.5–1.9)

 

<5 unit-years

24

1.0 (0.6–2.0)

 

5–25 unit-years

11

0.8 (0.4–1.6)

 

>25 unit-years

15

0.8 (0.4–1.7)

 

Females

 

 

No exposure

0

 

<5 unit-years

1

0.6 (0.1–5.3)

 

5–25 unit-years

0

 

>25 unit-years

0

Anttila et al., 1995

Biologically monitored workers in Finland

 

 

Entire period since first measurement

25

0.92 (0.59–1.35)

 

0–9 years

11

1.19 (0.59–2.13)

 

10–19 years

9

0.67 (0.30–1.26)

 

≥20 years

5

1.11 (0.36–2.58)

 

Mean personal U-TCA level

 

 

<100 µmol/L

16

1.02 (0.58–1.66)

 

100+ µmol/L

7

0.83 (0.33–1.71)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

All exposed factory workers

78

0.76 (0.60–0.95)

 

Duration of potential exposure (routine or intermittent)

 

 

<1 year

66

0.85 (0.65–1.13)

 

1–4 years

63

0.98 (0.74–1.30)

 

≥5 years

44

0.64 (0.46–0.89)

Blair et al., 1998

Aircraft-maintenance workers in Utah

 

 

Males

 

 

No exposure

51

1.0 (0.7–1.6)

 

<5 unit-years

43

1.0 (0.6–1.6)

 

5–25 unit-years

23

0.9 (0.5–1.6)

 

>25 unit-years

38

1.1 (0.7–1.8)

 

Females

 

 

No exposure

2

0.4 (0.1–1.6)

 

<5 unit-years

2

0.6 (0.1–2.4)

 

5–25 unit-years

11

0.6 (0.1–4.7)

 

>25 unit-years

2

0.4 (0.1–1.8)

Morgan et al., 1998

Aerospace workers in Arizona

 

 

Entire trichloroethylene-exposed cohort

97

1.10 (0.89–1.34)

 

Cumulative

 

 

Low

45

1.49 (1.09–1.99)

 

High

52

0.90 (0.67–1.20)

 

Peak: medium and high vs low and no exposure

64

1.07 (0.82–1.40)

Greenland et al., 1994

White male transformer-assembly workers, ever exposed

NA

1.01 (0.69–1.47)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Wilcosky et al., 1984

Rubber-industry workers in Ohio

 

 

Cumulative exposure of more than 1 year

 

 

White males

11

0.64

Tetrachloroethylene and Dry-cleaning Solvents

Cohort Studies—Mortality

Ruder et al., 2001

US dry-cleaning workers in four labor unions

 

 

Males, nonwhite

25

1.52 (1.05–2.39)

 

Females, nonwhite

16

1.88 (1.07–3.05)

 

Exposed to tetrachloroethylene only

19

1.17 (0.71–1.83)

 

Exposed to tetrachloroethylene and other dry-cleaning solvents

46

1.46 (1.07–1.95)

Boice et al., 1999

Aircraft manufacturing workers in California

 

 

All exposed factory workers

46

1.08 (0.79–1.44)

 

Duration of potential exposure (routine or intermittent)

 

 

<1 year

33

1.15 (0.80–1.66)

 

1–4 years

51

1.09 (0.80–1.48)

 

≥5 years

36

0.71 (0.49–1.02)

Anttila et al., 1995

Biologically monitored workers in Finland

5

1.92 (0.62–4.48)

Blair et al., 1990

Dry-cleaning union members in Missouri

47

1.3 (0.9–1.7)

Wilcosky et al., 1984

Rubber-industry workers in Ohio

 

 

Cumulative exposure of more than 1 year

 

 

White males

2

0.26

Case-Control Studies

Pohlabeln et al., 2000

Occupational exposure among nonsmoking females in Europe

 

 

Laundry and dry cleaners for at least 6 months

19

1.83 (0.98–3.40)

Paulu et al., 1999

Residents in upper Cape Cod

 

 

>75th percentile tetrachloroethylene-water exposure

 

 

0-year latent period

11

1.8 (0.8–3.9)

 

5-year latent period

6

1.7 (0.6–4.5)

 

7-year latent period

5

1.6 (0.5–4.4)

 

9-year latent period

4

1.8 (0.5–6.0)

 

>90th percentile tetrachloroethylene-water exposure

 

 

0-year latent period

5

3.7 (1.0–11.7)

 

5-year latent period

3

3.3 (0.6–13.4)

 

7-year latent period

3

6.2 (1.1–31.6)

 

9-year latent period

3

19.3 (2.5–141.7)

Brownson et al., 1993

Occupational exposure among females in Missouri—Dry-cleaning

 

 

Lifetime nonsmokers

23

2.1 (1.2–3.7)

 

Exposure range

 

 

Low: ≤1.125 years

NA

0.6 (0.2–1.7)

 

High: ≥1.125 years

NA

2.9 (1.5–5.4)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Methylene Chloride

Cohort Studies—Mortality

Hearne et al., 1999

Male cellulose triacetate photographic-film base workers in Kodak Park, employed >1 year

 

 

Methylene chloride cohort

27

0.75 (0.49–1.09)

 

Roll-coating cohort

 

 

New York state external control

28

0.82 (0.55–1.19)

 

Kodak Rochester external control

28

0.89 (0.59–1.29)

Tomensen et al., 1997

Male cellulose triacetate-fiber production workers in the UK, ever employed

 

 

Never exposed

1

0.16 (0–0.88)

 

All exposed

19

0.48 (0.29–0.75)

 

Cumulative exposure (nonsmokers)

 

 

<400 ppm-years

6

0.32

 

400–799 ppm-years

2

0.51

 

≥800 ppm-years

1

0.37

 

Unassigned exposure

10

0.68

Gibbs et al., 1996

Cellulose triacetate-fiber production workers in Maryland

 

 

Males

 

 

No exposure

6

0.59 (0.22–1.29)

 

Low exposure

20

0.78 (0.48–1.20)

 

High exposure

15

0.55 (0.31–0.91)

 

Females

 

 

No exposure

0

NA (0.0–4.92)

 

Low exposure

9

1.09 (0.50–2.07)

 

High exposure

2

2.29 (0.28–8.29)

Lanes et al., 1993

Cellulose triacetate-fiber production workers in South Carolina (cohort), employed >3 months

13

0.80 (0.43–1.37)

Benzene

Cohort Studies—Mortality

Yin et al., 1996a

Chinese factory workers, ever exposed

 

 

Total cohort

125

1.4 (1.0–2.0)

 

Males

109

1.5 (1.0–2.2)

 

Females

16

1.0 (0.4–2.9)

Hayes et al., 1996

Chinese factory workers

 

 

Cumulative exposure

 

 

No exposure

41

1.0

 

<10 ppm-years

10

1.2

 

10–39 ppm-years

13

1.0

 

40–99 ppm-years

19

1.4

 

100–400 ppm-years

38

1.4

 

400+ ppm-years

41

1.7

 

 

 

p-trend=0.01

Greenland et al., 1994

White male transformer-assembly workers, ever exposed

NA

0.58 (0.31–1.07)

Wilcosky et al., 1984

Rubber-industry workers in Ohio

 

 

Cumulative exposure of more than 1 year

 

 

White males

23

0.69

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Toluene

Cohort Studies—Mortality

Wiebelt and Becker, 1999

Male German rotogravure printing-plant workers, employed >1 year

 

 

Total cohort

44

1.23 (0.81–1.92)

 

Printing-cylinder preparation workers

7

0.83 (0.20–2.77)

 

Printing/proof printing workers

25

1.30 (0.72–2.49)

 

Finishing workers

13

1.77 (0.77–4.39)

Svensson et al., 1990

Male rotogravure printing-plant workers in Sweden

 

 

Total cohort

16

1.76 (1.03–2.91)

 

≥5 year exposure, >10 year latency

9

1.26 (0.57–2.38)

Sorahan and Cathcart, 1989

Male chemical-factory workers

 

 

Low exposure to chlorinated toluenes

16

1.39 (0.80–2.27)a

 

High exposure to chlorinated toluenes

10

3.31 (1.59–6.09)a

 

Benzotrichloride

NA

1.36 (0.43–4.24)

 

Other chlorinated toluenes

NA

1.12 (0.30–4.22)

Wilcosky et al., 1984

White, male rubber-industry workers in Ohio

 

 

Cumulative exposure for more than 1 year

3

0.55

Phenol

Cohort Studies—Mortality

Kauppinen et al., 1993

Finnish woodworkers

 

 

Any exposure (>1 month)

5

4.04 (1.83–8.89)

 

Duration >5 years

6

3.08 (0.70–13.6)

Dosemeci et al., 1991

US white male industrial workers

 

 

Any exposure

146

1.1 (0.9–1.3)

 

Level of cumulative exposure

 

 

None

70

1.2

 

Low exposure

68

1.2

 

Medium exposure

60

1.1

 

High exposure

18

1.4

Wilcosky et al., 1984

White male rubber-industry workers in Ohio

 

 

Cumulative exposure for more than 1 year

 

 

White males

13

0.95

 

Black males

2

0.91

Other Organic Solvents

Cohort Studies—Mortality

Anttila et al., 1995

Biologically monitored workers in Finland

 

 

1,1,1-trichloroethane ever measured in urine

2

1.31 (0.16–4.71)

Wilcosky et al., 1984

Rubber-industry workers in Ohio

 

 

Cumulative exposure for more than 1 year

 

 

Specialty naphthas (white males)

43

0.70

 

Specialty naphthas (black males)

2

0.39

 

Ethanol (white males)

21

1.0

 

Xylenes (white males)

10

0.61

 

Ethyl acetate (white males)

6

0.84

 

Acetone (white males)

5

0.86

 

Isopropanol (white males)

27

0.64

 

Isopropanol (black males)

2

0.56

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Unspecified Mixtures of Organic Solvents

Cohort Studies—Incidence

Malker and Gemne, 1987

Swedish printing-industry workers, employed in 1960

 

 

Males

190

1.5 (1.2–1.7)b

 

Females

9

1.3

Engholm and England, 1982

Male members of the Swedish Painters Union

 

 

Years since entry into the union

 

 

≥0

81

1.28 (p<0.05)

 

≥5

75

1.24 (p<0.05)

 

≥10

74

1.31 (p<0.05)

 

≥15

66

1.28 (p<0.05)

 

≥20

58

1.26 (p<0.05)

 

≥25

51

1.32 (p<0.05)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Potential routine exposure to mixed solvents

221

0.88 (0.77–1.01)

Steenland and Palu, 1999

Members of US painters unions

 

 

Total cohort

1746

1.23 (1.17–1.29)

 

20 years since first union membership

1360

1.24 (1.18–1.31)

Greenland et al., 1994

White male transformer-assembly workers, ever exposed

NA

1.57 (1.08–2.27)

Acquavella et al., 1993

Metal-components manufacturing workers

 

 

Solvents, ever exposed

4

1.9 (0.5–4.9)

Walker et al., 1993

Shoe-manufacturing workers in Ohio, employed >1 month

 

 

Total cohort

99

1.47 (1.20–1.80)

 

Males

68

1.56 (1.22–1.99)

 

Females

31

1.30 (0.89–1.86)

Teta et al., 1992

Male ethanol and isopropanol production workers

 

 

South Charleston SC plant

 

 

All workers, ever employed

14

0.87 (0.5–1.5)

 

Workers in exposed unit ≥10 years

2

0.56

 

Texas City, TX plant

 

 

All workers, ever employed

8

1.10 (0.5–2.2)

 

Workers in exposed unit ≥10 years

1

NA

Matanoski et al., 1986

US painters and allied tradesmen

 

 

Total cohort

448

1.06 (0.96–1.16)

Stockwell and Matanoski, 1985

Male construction and maintenance painters in New York

 

Usual occupation of painter

51

2.75 (1.45–5.21)

Engholm and England, 1982

Male members of the Swedish Painters Union

 

 

Years since entry into the union:

 

 

≥0

124

1.27 (p<0.05)

 

≥5

118

1.25 (p<0.05)

 

≥10

114

1.28 (p<0.05)

 

≥15

103

1.27 (p<0.05)

 

≥20

92

1.26 (p<0.05)

 

≥25

80

1.27 (p<0.05)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Morgan et al., 1981

Male US paint or varnish manufacturing workers

 

 

Solvents excluding lacquer, 1+ years of exposure

51

1.14

Waxweiler et al., 1981

Male synthetic-chemical plant workers, ever employed

42

1.49 (1.08–2.03)a

Alderson and Rattan, 1980

Male workers at dewaxing plants in the UK, employed >1 year

 

 

Workers in isopropanol alcohol plant

2

0.78 (0.09–2.81)a

 

Workers in methyl ethyl ketone plant

1

0.17 (0.00–0.93)a

Englund, 1980

Male Swedish painters, ever certified or union member

124

1.27 (1.06–1.52)a

Case-Control Study

Brownson et al., 1993

Occupational exposure among females in Missouri—Printing industry

 

 

Lifetime nonsmokers

6

0.8 (0.3–2.0)

 

Exposure range

 

 

Low: ≤8 years

NA

0.6 (0.2–2.2)

 

High: >8 years

NA

1.3 (0.5–3.7)

NOTE: NA=not available; U-TCA=urinary metabolite of trichloroethylene.

a95% CI calculated by the committee with standard methods from the observed and expected numbers presented in the original study.

b99% confidence limits.

BONE CANCER

Epidemiologic Studies of Exposure to Organic Solvents and Bone Cancer

Blair and colleagues (1998) extended the followup of a cohort of aircraft-manufacturing workers (Spirtas et al., 1991) that used a detailed exposure-assessment method. An increased bone cancer risk was reported after adjustment for age, calendar time, and sex (RR=2.1, 95% CI =0.2–18.8) in workers exposed to trichloroethylene.

One study assessed whether an association existed between exposure to benzene and bone cancer. To evaluate the specific relationship between exposure to benzene and cancer risk, Wong (1987a) examined mortality among workers employed in seven chemical-manufacturing plants. An increased relative risk of bone cancer was found in the exposed group of workers (SMR=3.17, 95% CI=0.38–11.46). Wong (1987b) also estimated exposure to benzene in terms of 8-hour TWAs and peak levels of exposure and found that the relative risk of bone cancer increased with duration of exposure (SMR=6.63; one exposed case). Those results were inconclusive for an association because of the very small number of exposed cases, which resulted in highly variable risk estimates.

Fu and colleagues (1996) examined two historical cohorts of shoe workers in England and Florence, Italy, and used job titles to assess cancer mortality in relation to exposure to leather dusts and solvents. A slight increase in bone cancer was observed in the English cohort among those with probable solvent exposure (SMR=1.12, 95% CI=0.03–6.26), and no cases of bone cancer were reported in the Florence cohort. The committee reviewed the study by Nielsen and colleagues (1996) in which the risk of bone cancer in a cohort of lithographers was examined. Only one exposed case was observed (SIR=11.4, 95% CI=0.6–56.0).

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Summary and Conclusion

The relationship between exposure to organic solvents and bone cancer was reported in only four cohort studies, representing three different solvent exposures. Each study had low power, and exposure assessment relied primarily on job titles as surrogates of exposure. No case-control studies of bone cancer were identified. Studies with larger numbers of exposed cases (increased power) and more precise exposure assessment are needed for the committee to evaluate the relationship between solvent exposures and risk of bone cancer. Table 6.16 identifies the key studies and relevant data points reviewed by the committee in drawing its conclusion. Unless indicated in the table, the study populations include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to solvents under review and bone cancer.

TABLE 6.16 Selected Epidemiologic Studies—Bone Cancer and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Study—Mortality

Blair et al., 1998

Aircraft maintenance workers in Utah, employed in exposed area >1 year

5

2.1 (0.2–18.8)

Benzene

Cohort Study—Mortality

Wong, 1987a,b

Male Chemical Manufacturers Association workers

 

 

Continuous exposure

2

3.17 (0.38–11.46)

 

Duration of exposure

 

 

<5 years

1

2.63

 

5–14 years

1

6.63

 

≥15 years

0

Unspecified Mixtures of Organic Solvents

Cohort Study—Incidence

Nielsen et al., 1996

Danish lithographers, ever employed

1

11.4 (0.6–56.0)

Cohort Study—Mortality

Fu et al., 1996

Shoe-manufacturing workers

 

 

English cohort, employed in 1939

6

2.08 (0.76–4.52)

 

Probable solvent based on work area

1

1.12 (0.03–6.26)

 

Florence cohort, ever employed

0

0 (0.0–3.45)

SOFT TISSUE SARCOMA

Because of the lack of available studies on the relationship between exposure to organic solvents and soft tissue sarcomas (STS), a conclusion regarding association could not be drawn. Only one study (Serraino et al., 1992) identified by the committee analyzed the relationship between relevant exposures reviewed in this report (“benzene/solvents” and “dyes/paints”) and STS. Although the population-based case-control study observed an increased risk of STS among men exposed to “benzene/solvents” for more than 10 years (OR=2.2, 95% CI=0.9–5.5), the study was limited by the use of self-reported exposures. Additional studies are needed to support the relationship before a conclusion regarding association can be drawn.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

SKIN CANCER

Description of Case-Control Studies

Table 6.17 identifies the study characteristics of two papers from the Montreal multisite cancer case-control study on the association between exposure to specific organic solvents and solvent mixtures and the risk of melanoma. The study was designed so that in-person interviews were used mostly, and telephone interviews or self-administered questionnaires were limited to next of kin or hard-to-interview subjects. The interviews included a job-specific module to obtain detailed information on each job that a subject held in the entire working history, such as dates of employment, the employer’s activities and products, job tasks, and work environment. Using the job-history information, a team of industrial hygienists and chemists estimated exposures to about 300 of the most common occupational agents. The study population was used for exposure-specific and cancer-specific studies. Fritschi and Siemiatycki (1996a) evaluated the relationship between melanoma and exposure to 85 chemical substances, 13 occupations, and 20 industries. The same set of cases was studied by Gérin and colleagues (1998) in relation to occupational exposure to the hydrocarbons benzene, toluene, xylene, and styrene.

Epidemiologic Studies of Exposure to Organic Solvents

No increase in melanoma mortality was found among two cohorts of aircraft-maintenance workers (Blair et al., 1998; Boice et al., 1999), a cohort of workers monitored for a metabolite of trichloroethylene (Hansen et al., 2001), or a cohort of uranium-processing workers (Ritz, 1999). An increased risk of melanoma (for “any exposure,” OR=3.6, 95% CI=1.5–9.1) was found in a case-control study in Montreal (Fritschi and Siemiatycki, 1996a). Two studies of workers monitored for exposure to trichloroethylene found mixed results for nonmelanoma skin cancers (Axelson et al., 1994; Hansen et al., 2001). The studies were based on small numbers of exposed cases and did not control for exposure to sunlight, an important confounding variable.

Two studies of tetrachloroethylene-exposed workers, one in aircraft manufacturing (Boice et al., 1999) and one of dry-cleaning workers (Blair et al., 1990), showed no association between exposure and skin cancer risk. Boice and colleagues examined melanoma specifically, and Blair and colleagues looked at all skin cancers combined.

Several other solvents thought to have been used in the Gulf War—including methylene chloride, benzene, toluene, xylene, and phenol—were investigated in relation to melanoma. The committee identified only one relevant epidemiologic study for each substance. A cohort study of cellulose-fiber production workers showed an increased risk of melanoma mortality (SMR=1.94 95% CI=0.24–7.00) (Lanes et al., 1993). No increased risks were reported for melanoma and exposure to benzene, toluene, or xylene (Gérin et al., 1998) or for any type of skin cancer and exposure to phenol (Dosemeci et al., 1991).

Melanoma was not found to be associated with exposure to unspecified mixtures of solvents, as reported in several occupational studies (Anttila et al., 1995; Berlin et al., 1995; Boice et al., 1999; Bourguet et al., 1987; Fritschi and Siemiatycki, 1996a). The findings on risk of nonmelanoma skin cancer and unspecified mixtures of solvents were inconsistent but tended to be negative. Studies that found positive associations were variable and based on few exposed cases, and sunlight exposure was not controlled for.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.17 Description of Case-Control Studies of Melanoma Skin Cancers and Exposure to Organic Solvents

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Fritschi and Siemiatycki, 1996a

Male cases and controls, age 35–70 years, diagnosed in 19 large Montreal-area hospitals in 1979–1985 and histologically confirmed for one of 19 anatomic cancer sites; age-matched, population-based controls were also chosen from electoral lists and RDD (see also Gérin et al., 1998)

103 melanoma

1,066 subjects for each site, consisting of

533 population controls and

533 randomly selected subjects from the eligible cancer control group

Trichloroethylene

Solvents

In-person interviews with segments on work histories (job titles); exposures attributed by a team of chemists and industrial hygienists

Unconditional logistic regression

Age, years of schooling, ethnicity

Response rates: 83% of all cases, 71% of population controls

Gérin et al., 1998

Male cases and controls, age 35–70 years, diagnosed in 19 large Montreal-area hospitals in 1979–1985 and histologically confirmed for one of 19 anatomic cancer sites; age-matched, population-based controls were also chosen from electoral lists and RDD (see also Fritschi and Siemiatycki, 1996a)

103 melanoma

1,066 subjects for each site, consisting of

533 population controls and

533 randomly selected subjects from the eligible cancer control group

Benzene

Toluene

Xylene

In-person interviews with segments on work histories (job titles); exposures attributed by a team of chemists and industrial hygienists

Unconditional logistic regression

Age, family income, ethnicity, cigarette smoking, respondent status

Response rates: 82% of all cases, 71% of population controls

NOTE: RDD=random-digit dialing

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Summary and Conclusion

Several studies of specific and mixed solvents examined the role of exposure and the risk of skin cancer, specifically melanoma and nonmelanoma. However, small numbers of exposed cases and the lack of validated exposure assessment were limitations of the studies. Almost all studies on specific organic solvents and unspecified mixtures of organic solvents found no association between exposure and incidence or mortality from melanoma, nonmelanoma skin cancer, or skin cancer in general. Several studies used biologic monitoring to assess exposure (e.g., Anttila et al., 1995; Axelson et al., 1994; Hansen et al., 2001), but the inability of most studies to control for well-established risk factors of skin cancer—such as age, ethnicity, geography, presence of nevi, and time spent in the sun—limits the validity of their findings. Tables 6.1819 identify the key studies and relevant data points reviewed by the committee in drawing its conclusions. Unless indicated in the tables, the study populations include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure solvents under review and melanoma or nonmelanoma skin cancer.

TABLE 6.18 Selected Epidemiologic Studies—Melanoma Skin Cancers and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Study—Incidence

Hansen et al., 2001

Biologically monitored male Danish workers

2

0.9 (0.1–3.4)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers, ever exposed

2

0.46 (0.06–1.67)

Ritz, 1999

White male uranium-processing workers, ever exposed

 

 

Melanoma and nonmelanoma

4

0.64 (0.17–1.63)

Blair et al., 1998

Aircraft-maintenance workers, ever exposed

9

1.0 (0.3–3.1)

Case-Control Study

Fritschi and Siemiatycki, 1996a

Male residents of Montreal, Canada

 

Any exposure

8

3.6 (1.5–9.1)

 

Substantial exposure

4

3.4 (1.0–12.3)

Tetrachloroethylene and Dry-cleaning Solvents

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers, ever exposed

2

0.95 (0.12–3.43)

Blair et al., 1990

Dry-cleaning union members in Missouri, ever employed

 

 

Melanoma and nonmelanoma

2

0.8 (0.1–2.8)

Other Specific Organic Solvents

Cohort Studies—Mortality

Lanes et al., 1993

Cellulose-fiber production workers in South Carolina, exposed to methylene chloride

2

1.94 (0.24–7.00)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Dosemeci et al., 1991

Male US industrial workers, ever exposed to phenol

 

 

Melanoma and nonmelanoma

7

0.9 (0.4–1.8)

Case-Control Study

Gérin et al., 1998

Male residents of Montreal, Canada, any exposure

 

 

Benzene

11

0.6 (0.3–1.2)

 

Toluene

5

0.4 (0.1–0.9)

 

Xylene

3

0.3 (0.1–0.8)

Unspecified Mixtures of Organic Solvents

Cohort Studies—Incidence

Anttila et al., 1995

Workers in Finland biologically monitored for halogenated hydrocarbons

5

0.71 (0.23–1.66)

Berlin et al., 1995

Patients with solvent-related disorders, ever exposed

3

0.7 (0.1–2.0)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers, ever exposed

 

 

Mixed solvents

10

0.87 (0.42–1.60)

Morgan et al., 1981

Male paint and coatings manufacturing workers, ever employed

 

 

Melanoma and nonmelanoma

4

1.48

Case-Control Study

Fritschi and Siemiatycki, 1996a

Male residents of Montreal, Canada, ever exposed

 

Solvents

33

0.8 (0.5–1.3)

TABLE 6.19 Selected Epidemiologic Studies—Nonmelanoma Skin Cancers and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Study—Incidence

Hansen et al., 2001

Biologically monitored male Danish workers

15

1.0 (0.6–1.6)

Cohort Studies—Mortality

Ritz, 1999

White male uranium-processing workers, ever exposed

 

 

Melanoma and nonmelanoma

4

0.64 (0.17–1.63)

Axelson et al., 1994

Biologically monitored male Swedish workers

 

 

Nonmelanoma

8

2.36 (1.02–4.65)

Tetrachloroethylene and Dry-cleaning Solvents

Cohort Study—Mortality

Blair et al., 1990

Dry-cleaning union members in Missouri, ever employed

 

 

Melanoma and nonmelanoma

2

0.8 (0.1–2.8)

Other Specific Organic Solvents

Cohort Study—Mortality

Dosemeci et al., 1991

Male US industrial workers, ever exposed to phenol

 

 

Melanoma and nonmelanoma

7

0.9 (0.4–1.8)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Unspecified Mixtures of Organic Solvents

Cohort Studies—Incidence

Anttila et al., 1995

Workers in Finland biologically monitored for halogenated hydrocarbons

2

0.46 (0.06–1.67)

Berlin et al., 1995

Patients with solvent-related disorders, ever exposed

4

1.5 (0.4–4.0)

Bourguet et al., 1987

Male tire and rubber manufacturing workers, ever exposed

 

 

Low solvent exposure

15

0.6

 

Medium solvent exposure

7

1.1

 

High solvent exposure

34

1.1

Cohort Study—Mortality

Morgan et al., 1981

Male paint and coatings manufacturing workers, ever employed

 

 

Melanoma and nonmelanoma

4

1.48

BREAST CANCER

Description of Case-Control Studies

All breast cancer case-control studies were population-based and are identified in Table 6.20 below. Aschengrau and colleagues (1998) evaluated the relationship between the risk of breast cancer and exposure to tetrachloroethylene in drinking water, which was estimated on the basis of an algorithm that accounted for residential history, water flow, and pipe characteristics, as established by Webler and Brown (1993). Three other case-control studies evaluated breast cancer risk and occupational exposure (Band et al., 2000; Hansen, 1999; Petralia et al., 1999). Each of the studies ascertained exposure differently: through use of occupational titles (Band et al., 2000), through linkage of pension-fund occupational-history information with solvent use in industries (Hansen, 1999), and through interviews with subjects to obtain occupational histories, which were linked with job-exposure matrixes to assign cumulative exposure measures (Petralia et al., 1999). Potential confounding variables were handled adequately in three of the four studies (Aschengrau et al., 1998; Band et al., 2000; Petralia et al., 1999).

Epidemiologic Studies of Exposure to Organic Solvents and Breast Cancer

Risk of breast cancer was not increased in two Scandinavian studies of biologically monitored workers exposed to trichloroethylene (Anttila et al., 1995; Hansen et al., 2001). Among women exposed to trichloroethylene as aircraft-maintenance workers, Blair and colleagues (1998) found a risk of breast cancer associated with any exposure to trichloroethylene (SMR=1.8, 95% CI=0.9–3.3) and more than a 3-fold risk associated with continuous low exposure to trichloroethylene (SMR=3.4, 95% CI=1.4–8.0). Other cohort studies of aircraft workers in which trichloroethylene was considered a predominant solvent did not report increased mortality rates from breast cancer (male and female combined) (Boice et al., 1999; Morgan et al., 1998).

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.20 Description of Case-Control Studies of Breast Cancer and Exposure to Organic Solvents

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Aschengrau et al., 1998

Female cases reported to the Massachusetts Cancer Registry, diagnosed in 1983–1986 among residents of five upper Cape Cod towns; living controls were selected from the records of HCFA and through RDD; deceased controls identified by the state Department of Vital Statistics and Research files

258

686

Tetrachloroethylene

Relative delivered dose estimated in model accounting for location and years of residence, water flow, pipe characteristics

Multiple logistic regression

Age at diagnosis, vital status, family and personal history of breast cancer or disease, age at first birth, occupational exposure to solvents

Response rates: 79% of cases, 76% of HCFA controls, 74% of RDD controls, 79% of next of kin of deceased controls

Petralia et al., 1999

Female cases, age 40 years or more, identified through major hospitals in two New York counties in 1986–1991, with histologic confirmation; controls randomly selected from lists of the NY State Department of Motor Vehicles, matched for age and county

301

316

Benzene

In-person interviews to assess lifetime occupational history; occupations and industries coded; assigned potential exposures to polycyclic aromatic hydrocarbons through use of a job-exposure matrix

Unconditional logistic regression

Age, years of education, age at first birth, age at menarche, history of benign breast disease, family breast cancer history, Quetelet index, months of lactation

Participation rates: 66% of cases, 62% of controls

Hansen, 1999

Female cases, identified through the Danish Cancer Registry, born in 1934–1969 with diagnosis in 1970–1989; controls randomly selected from the central population register, matched for year of birth and sex

7,802

7,802

Industries with extensive solvent use

Past employment determined through linkage to the national pension fund files; occupations from five industrial groupings (except administrative jobs) classified as exposed to solvents

Conditional logistic regression

Age, social class, age at first child, number of children

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Band et al., 2000

Female cases, under 75 years old, identified through the British Columbia Cancer Registry as Canadian citizens residing in British Columbia and diagnosed in 1988–1989, with histologic confirmation; controls selected from voter list, matched for age

995

1020

Laundering and dry-cleaning (occupation and industry)

Mailed questionnaire to assess lifetime job history with occupational and industry coding

Conditional logistic regression

Ethnicity, age at menarche and menopause, smoking, marital status, education, alcohol consumption, other medical factors

Response rates: 74.7% of cases, 76.1% of controls

NOTE: HCFA=Health Care Financing Administration; RDD=random-digit dialing..

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Some epidemiologic studies have either specifically addressed exposure to tetrachloroethylene or examined the effect of dry-cleaning work in relation to breast cancer risk. Aschengrau and colleagues (1998) conducted a case-control study in the Cape Cod area where drinking water was contaminated with tetrachloroethylene. Increased risks of breast cancer were found with longer latency. Excluding exposures occurring within the 9 years before diagnosis, the adjusted relative risks of breast cancer increased with increasing exposure (exposure greater than the 90th percentile: OR=7.8, 95% CI=0.9–167.0). A study of aircraft-manufacturing workers showed a slight increase in mortality from breast cancer with potential routine exposure to tetrachloroethylene (SMR=1.16, 95% CI=0.32–2.97) (Boice et al., 1999). A cohort of dry-cleaning union members showed no increase in breast cancer mortality (SMR=0.91, 95% CI=0.55–1.40) or in workers exposed only to tetrachloroethylene (SMR=0.78, 95% CI=0.28–1.69) (Ruder et al., 2001).

Band and colleagues (2000) examined the potential risk of premenopausal and postmenopausal breast cancer in multiple occupations in British Columbia. Increased risk of breast cancer was observed in women reporting any or usual work in laundry and dry-cleaning. Exposure to tetrachloroethylene and other solvents was based on an occupational title of “laundering/dry-cleaning” for premenopausal women and dry-cleaning occupational and industry titles for postmenopausal women. Among numerous positive associations, postmenopausal women experienced an almost 5-fold risk if their usual occupation was laundry and dry-cleaning (OR=4.85, 95% CI=1.26–18.7). The other cohort studies of dry cleaners did not show positive associations between this employment and breast cancer risk in women (Blair et al., 1990).

The association between exposure to benzene and the risk of breast cancer was assessed in two cohort studies and one case-control study. Petralia and colleagues (1999) reported an increased adjusted risk of premenopausal breast cancer with exposure to benzene and that risk increased with probability and duration of exposure (duration at least 4 years: OR=3.38, 95% CI=1.25–9.17). The large cohort study of Chinese benzene-exposed workers did not show an increase in breast cancer mortality (RR=0.9, 95% CI=0.3–3.2) (Yin et al., 1996a). The Danish cohort study, in which exposure was assessed using pension-fund records of job history, showed an increased incidence of breast cancer in men (OR=2.2, 95% CI=1.4–3.6) (Hansen, 2000).

The cohort of aircraft-maintenance workers studied by Blair and colleagues experienced an increased risk of breast cancer with exposure to methylene chloride (RR=3.0, 95% CI=1.0–8.8). The study also produced increased relative risks of breast cancer with exposure to several specific solvents; positive associations with 1,1,1-trichloroethane, acetone, isopropyl alcohol, toluene, and methyl ethyl ketone were found.

Employees of a cellulose-fiber production plant with heavy methylene chloride use did not experience a rate of breast cancer higher than that in the local county population (SMR=0.54, 95% CI=0.11–1.57) (Lanes et al., 1993).

Several studies examined the potential relationship between breast cancer risk and exposure to mixtures of solvents (Anttila et al., 1995, 1998; Berlin et al., 1995; Blair et al., 1998; Cocco et al., 1998; Hansen, 1999, Shannon et al., 1988; Weiderpass et al., 1999). The studies characterized the exposure in general terms, such as “organic solvents” or “mixed solvents.” In two instances (Anttila et al., 1995, 1998), specific constituent solvents were mentioned, but separate analyses were not performed. Hansen (1999) found increased risks of breast cancer (predominantly premenopausal) associated with occupations and industries with heavy solvent use (over 10 years of employment: OR=1.31, 95% CI=1.01–1.75). In a cohort study of lamp manufacturers, Shannon and colleagues (1998) found a 2-fold risk of breast cancer in solvent-

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

exposed workers in coiling and wire drawing. An association with any exposure to solvents was found among aircraft-maintenance workers (SMR=1.6, 95% CI=0.9–2.8) (Blair et al., 1998). No associations were found in studies of workers exposed to aromatic or halogenated solvents (Anttila et al., 1995, 1998), in patients with solvent-related disorders (Berlin et al., 1995), and in aircraft-manufacturing workers (Garabrant et al., 1988).

Summary and Conclusion

In most occupational settings, multiple solvent exposures occurred, so exposures to specific solvents may be highly correlated. Because many studies used occupational titles as exposure surrogates, the ability to assess an association between specific solvents and breast cancer risk was compromised.

A number of studies assessed breast cancer risk and solvent exposure in general, and others provided exposure estimates for specific individual solvents, such as trichloroethylene, tetrachloroethylene, dry-cleaning solvents, benzene, and methylene chloride. The evidence was limited by nonspecific exposure assessments and a reliance on mortality from breast cancer. Nondifferential misclassification of exposure, poor control for confounding, and low statistical power due to small numbers were additional limitations. Table 6.21 identifies the key studies and relevant data points reviewed by the committee in drawing its conclusion. Unless indicated in the tables, the study populations include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to solvents under review and breast cancer.

TABLE 6.21 Selected Epidemiologic Studies—Breast Cancer and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Study—Incidence

Hansen et al., 2001

Biologically monitored workers in Denmark

 

 

Females, ever exposed

4

0.9 (0.2–2.3)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Potential routine exposure

7

1.31 (0.53–2.69)

Blair et al., 1998

Aircraft-maintenance workers in Utah (females)

 

 

Any exposure

20

1.8 (0.9–3.3)

 

Low, continuous

8

3.4 (1.4–8.0)

 

Frequent peaks

10

1.4 (0.7–3.2)

Morgan et al., 1998

Aerospace workers in Arizona

 

Ever exposed

16

0.75 (0.43–1.22)

 

Low exposure

11

1.03 (0.51–1.84)

 

High exposure

5

0.47 (0.15–1.11)

Anttila et al., 1995

Finnish workers biologically monitored for exposure to halogenated hydrocarbons

34

0.85 (0.59–1.18)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Tetrachloroethylene and Dry-cleaning Solvents

Cohort Studies—Mortality

Ruder et al., 2001

US dry-cleaning workers

 

 

Dry-cleaning, employed >1 year

20

0.91 (0.55–1.40)

 

Tetrachloroethylene only

6

0.78 (0.28–1.69)

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Potential routine exposure

4

1.16 (0.32–2.97)

Blair et al., 1990

Dry-cleaning union members in Missouri (females)

36

1.0 (0.7–1.4)

Case-Control Studies

Band et al., 2000

Female cases from the British Columbia Cancer Registry

 

 

Premenopausal

 

 

Laundering or dry-cleaning (occupation)

 

 

Usual exposure

1

 

Ever exposed

4

1.77 (0.41–7.72)

 

Postmenopausal

 

 

Laundry or dry-cleaning (occupation)

 

 

Usual exposure

8

4.85 (1.26–18.7)

 

Ever exposed

8

1.33 (0.55–3.19)

 

Laundering and dry-cleaning (industry)

 

 

Usual exposure

9

5.24 (1.41–19.5)

 

Ever exposed

12

1.42 (0.68–2.99)

 

Power laundries or dry cleaners (industry)

 

 

Usual exposure

9

2.00 (0.78–5.13)

 

Ever exposed

21

1.67 (0.89–3.13)

Aschengrau et al., 1998

Residents of upper Cape Cod, MA

 

Postmenopausal women

 

 

75th percentile (9-year latency)

NA

3.4 (0.7–19.1)

 

90th percentile (9-year latency)

NA

7.8 (0.9–167.0)

Benzene

Cohort Study—Incidence

Hansen, 2000

Male members of the national pension fund in Denmark

 

 

No lag time

19

2.2 (1.4–3.6)

 

>10 years lag

12

2.5 (1.3–4.5)

Cohort Study—Mortality

Yin et al., 1996a

Chinese factory workers (females), ever exposed

8

0.9 (0.3–3.2)

Case-Control Study

Petralia et al., 1999

Female residents of New York state

 

 

Any exposure

56

1.91 (1.18–3.08)

 

Duration <4 years

8

0.80 (0.30–2.16)

 

Duration ≥4 years

16

3.38 (1.25–9.17)

 

Low probability

8

1.22 (0.42–3.56)

 

Medium or high probability

16

2.14 (0.89–5.12)

 

Low intensity

16

2.38 (0.97–5.87)

 

Medium or high intensity

8

1.07 (0.37–3.07)

 

Low cumulative

13

1.43 (0.59–3.47)

 

Medium or high cumulative

11

2.21 (0.77–6.36)

 

10 to 19 year latency

5

1.23 (0.34–4.46)

 

≥20 year latency

16

2.09 (0.85–5.14)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Methylene Chloride

Cohort Studies—Mortality

Blair et al., 1998

Aircraft-maintenance workers in Utah, ever exposed (females)

4

3.0 (1.0–8.8)

Lanes et al., 1993

Cellulose-fiber production-plant workers in South Carolina, employed >3 months

3

0.54 (0.11–1.57)

Other Specific Organic Solvents

Cohort Study—Mortality

Blair et al., 1998

Aircraft-maintenance workers in Utah, employed >1 year (females)

 

 

1,1,1-Trichloroethane

3

3.3 (1.0–11.2)

 

Acetone

7

1.9 (0.8–4.6)

 

Isopropyl alcohol

8

3.7 (1.6–8.4)

 

Methyl ethyl ketone

8

2.1 (0.9–4.7)

 

Stoddard solvent

15

1.2 (0.6–2.4)

 

Toluene

10

2.0 (0.9–4.2)

Unspecified Mixtures of Organic Solvents

Cohort Studies—Incidence

Anttila et al., 1998

Finnish workers biologically monitored for exposure to aromatic hydrocarbons

15

0.79 (0.44–1.30)

 

Latency 0–9 years

8

0.61 (0.26–1.21)

 

Latency 10+ years

7

1.18 (0.47–2.42)

Anttila et al., 1995

Finnish workers biologically monitored for exposure to halogenated hydrocarbons

34

0.85 (0.6–1.2)

 

Latency 0–9 years

12

0.84 (0.4–1.5)

 

Latency 10+ years

22

0.85 (0.5–1.3)

Berlin et al., 1995

Swedish patients with solvent-related disorders (females)

3

1.1 (0.2–3.2)

Garabrant et al., 1988

Aircraft-manufacturing workers in California, employed >4 years

16

0.91 (0.52–1.48)

Shannon et al., 1988

Lamp manufacturing workers, employed >6 months

 

Coiling or wire drawing

8

2.04 (0.88–4.02)

Cohort Study—Mortality

Blair et al., 1998

Aircraft-maintenance workers in Utah (females)

 

 

Organic solvents, ever exposed

28

1.6 (0.9–2.8)

Case-Control Studies

Hansen, 1999

Danish women employed in industries with heavy solvent use

 

 

Employed ≥1 year, no lag

743

1.27 (1.13–1.43)

 

Employed ≥1 year, ≥15 years lag

472

1.43 (1.24–1.67)

 

Employed >10 years, no lag

113

1.31 (1.01–1.75)

 

Employed >10 years, ≥15 years lag

97

1.97 (1.39–2.79)

Cocco et al. 1998

Male breast cancer cases from National Mortality Follow-Back Survey exposed to organic solvents

 

 

Low probability

26

0.8 (0.5–1.3)

 

Medium probability

15

0.9 (0.5–1.6)

 

High probability

1

0.5 (0.1–4.3)

 

Low intensity

18

0.8 (0.4–1.4)

 

Medium intensity

20

0.8 (0.5–1.5)

 

High intensity

4

0.7 (0.2–1.2)

NOTE: NA=not available.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

FEMALE REPRODUCTIVE CANCERS

Epidemiologic Studies of Exposure to Organic Solvents and Cervical Cancer

Anttila and colleagues (1995) conducted a followup study of cancer incidence among Finnish workers who were biologically monitored for exposure to trichloroethylene and other halogenated hydrocarbons. An increased risk of cervical cancer was observed among women with any exposure to trichloroethylene (SIR=2.42, 95% CI=1.05–4.77), and the relative risks increased with increasing exposure (RRlow=1.86, 95% CI=0.38–5.45; RRhigh=4.35, 95% CI=1.41–10.1). A similar biomonitoring study in Denmark showed a 3.8-fold risk of cervical cancer (SIR=3.8, 95% CI=1.0–9.8) (Hansen et al., 2001), but no exposure-response pattern was reported.

A retrospective cohort study by Blair and colleagues (1998) of US civilians employed in aircraft maintenance showed increased mortality from cervical cancer (SMR=3.0, 95% CI=0.5–6.5); the relative risk was increased among women with high cumulative exposure (SMR=3.0, 95% CI=0.8–11.7). Two studies of aerospace and aircraft-manufacturing workers (Boice et al., 1999; Morgan et al., 1998) found no deaths from cervical cancer.

Only two studies specifically examined the association between exposure to tetrachloroethylene and cervical cancer. However, given that tetrachloroethylene is often used in dry-cleaning work, studies on laundry and dry-cleaning workers were also reviewed.

Boice and colleagues (1999) did not observe any cases of cervical cancer among women exposed to tetrachloroethylene. The biomonitoring study of tetrachloroethylene-exposed workers by Anttila and colleagues (1995) found an increased risk of cervical cancer (SIR=3.20, 95% CI =0.39–11.6).

Ruder and colleagues (2001) found an increased risk of cervical cancer (SMR=1.95, 95% CI=1.00–3.40) among members of a dry-cleaning union representing four areas in the United States, and Blair and colleagues (1990) estimated an SMR of 1.7 (95% CI=1.0–2.0) in a study of dry-cleaning union members in Missouri. Ruder and colleagues (2001) found that the risk of cervical cancer increased further among women who were exposed for more than 5 years (SMR=2.78 for less than 20 years of latency and 2.40 for 20 years or more of latency), but Blair and colleagues (1990) found no increases in risk with increasing exposure.

The available evidence concerning an association between exposure to methylene chloride and risk of cervical cancer was sparse. Gibbs and colleagues (1996) examined the risk of cervical cancer among women employed in cellulose-fiber production. They found an increased risk of cervical cancer among workers with low or high exposure (SMRlow=2.96, 95% CI=0.96–6.92; SMRhigh=5.40, 95% CI=0.14–30.1), but it was also increased in workers with no measured exposure to methylene chloride (SMR=7.02, 95% CI=0.18–39.1). Shannon and colleagues (1988) studied lamp manufacturing workers who were exposed to methylene chloride and other solvents and substances used during the manufacturing process, particularly coiling and wire drawing, and found no increased risk of cervical cancer.

The study of aircraft-manufacturing workers by Boice and colleagues (1999) did not show any cases of cervical cancer in women exposed to mixtures of solvents. In 1995, Berlin and colleagues examined cancer incidence and mortality patterns among patients with suspected solvent-related disorders and found that the incidence of cervical cancer was high (SIR=3.7, 95% CI=2.2–6.2).

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Summary and Conclusion

A limited number of papers were found that reported the risk of cervical cancer in connection with exposure to specific solvents. There were no case-control studies, and most of the cohort studies had very few cases. The studies of cervical cancer and specific solvent exposures did not provide evidence of an association between most of the specific solvents or solvent mixtures except for trichloroethylene.

For exposure to trichloroethylene and cervical cancer, three cohort studies showed an increased risk of cervical cancer with exposure to trichloroethylene, and two other studies did not have sufficient numbers or followup to find any deaths. An exposure-response relationship for the highest exposure was reported in two of the biologic monitoring studies (Anttila et al., 1995; Blair et al., 1998). Some committee members believed that the evidence of an association between cervical cancer and exposure to trichloroethylene should be classified as limited/suggestive. However, some committee members were concerned about confounding by socioeconomic status and the increased risk of exposure to the human papilloma virus (HPV), which is associated with the development of cervical cancer (NCI, 2002). The studies compared the risk of cervical cancer among unskilled workers of low socioeconomic status in Scandinavian countries with that of the general population. The positive associations could have been attributed to the lack of control for socioeconomic status or HPV infection. Moreover, no trends were seen with duration of employment or cumulative exposure in the other biologically monitored study of Danish workers (Hansen et al., 2001). There was also a concern that the numbers of studies and exposed cases were too small to support a conclusion that the evidence was limited/suggestive. In addition, three studies (Blair et al., 1998; Boice et al., 1999; Morgan et al., 1998) had no exposed cases, although cases of cervical cancer were expected. Thus, some committee members concluded that the evidence was inadequate/insufficient to determine whether an association exists.

As a result, after extensive discussion, the committee could not reach a consensus as to whether the evidence was limited/suggestive of an association or was inadequate/insufficient to determine whether an association exists between cervical cancer and exposure to trichloroethylene. Future studies that control for socioeconomic status are needed to determine whether there is an association between exposure to trichloroethylene and the risk of cervical cancer.

For exposure to tetrachloroethylene and dry-cleaning solvents, although several studies were positive, exposure-response patterns were absent and limited the strength of the evidence. Only one study on the exposure to unspecified mixtures of organic solvents and risk of cervical cancer was identified. Although the finding was particularly strong, no exposure-relationship pattern was reported. Additional corroborating studies are needed before a determination can be made that an association exists between exposure to solvent mixtures and cervical cancer. Table 6.22 identifies the studies reviewed by the committee in making its conclusion regarding association.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to solvents under review, other than trichloroethylene, and cervical cancer.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.22 Selected Epidemiologic Studies—Cervical Cancer and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Studies—Incidence

Hansen et al., 2001

Biologically monitored workers in Denmark

4

3.8 (1.0–9.8)

Anttila et al., 1995

Biologically monitored workers in Finland

 

 

Whole period of followup (mean individual urinew level)

8

2.42 (1.05–4.77)

 

<100 µmol/L

3

1.86 (0.38–5.45)

 

100+ µmol/L

5

4.35 (1.41–10.1)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California, potential routine exposure

0

Blair et al., 1998

Aircraft-maintenance workers in Utah, ever exposed

5

1.8 (0.5–6.5)

 

<5 unit-years

1

0.9 (0.1–8.3)

 

5–25 unit-years

0

 

>25 unit-years

4

3.0 (0.8–11.7)

Morgan et al., 1998

Aerospace workers in Arizona, employed >6 months

0

Tetrachloroethylene and Dry-cleaning Solvents

Cohort Study—Incidence

Anttila et al., 1995

Biologically monitored workers in Finland

2

3.20 (0.39–11.6)

Cohort Studies—Mortality

Ruder et al., 2001

US dry-cleaning workers

 

 

Employed >1 year

12

1.95 (1.00–3.40)

 

Employed 5+ years, <20 years of latency

4

2.78 (0.75–7.71)

 

Employed 5+ years, ≥20 years of latency

3

2.40 (0.48–7.86)

Boice et al., 1999

Aircraft-manufacturing workers in California, potential routine exposure

0

Blair et al., 1990

Dry-cleaning union members in Missouri

21

1.7 (1.0–2.0)

Methylene Chloride

Cohort Studies—Mortality

Gibbs et al., 1996

Cellulose triacetate-fiber workers, employed >3 months

 

 

Methylene chloride, no exposure

1

7.02 (0.18–39.1)

 

Methylene chloride, low probability

5

2.96 (0.96–6.92)

 

Methylene chloride, high probability

1

5.40 (0.14–30.1)

Shannon et al., 1988

Lamp-manufacturing workers (primary exposure to methylene chloride), employed >6 months

1

1.05 (0.03–5.86)

Unspecified Mixtures of Organic Solvents

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Mixed solvents, potential routine exposure

0

Berlin et al., 1995

Swedish patients with solvent-related disorders

14

3.7 (2.2–6.2)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Epidemiologic Studies of Exposure to Organic Solvents and Ovarian and Uterine Cancer

There was a paucity of studies regarding exposure to specific solvents and ovarian or uterine cancer. No studies showed meaningful increases in the risk of uterine or ovarian cancer in relation to exposure to trichloroethylene. The number of exposed subjects was extremely small in these studies, so the data are not informative in drawing a conclusion regarding association. There were no reports of an association between risk of uterine or ovarian cancer and exposure to methylene chloride or of an association between ovarian cancer and exposure to unspecified mixtures of organic solvents (Boice et al., 1999).

Summary and Conclusion

A limited body of evidence was available for the committee to review concerning specific and unspecified solvent exposure and risk of ovarian or uterine cancer. Very few studies had sufficient power to permit meaningful analyses. Tables 6.23 and 6.24 identify the key studies reviewed by the committee in drawing its conclusions regarding various solvent exposures and ovarian or uterine cancer.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to solvents under review and ovarian or uterine cancer.

TABLE 6.23 Selected Epidemiologic Studies—Ovarian Cancer and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Study—Incidence

Hansen et al., 2001

Biologically monitored workers in Denmark

2

2.1 (0.2–7.6)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California, potential routine exposure

1

0.58 (0.01–3.22)

Morgan et al., 1988

Aerospace workers in Arizona, employed >6 months

 

 

Ever exposed

8

1.21 (0.52–2.38)

 

Low exposure

2

0.61 (0.07–2.21)

 

High exposure

6

1.79 (0.66–3.88)

Tetrachloroethylene and Dry-cleaning Solvents

Cohort Study—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California, potential routine exposure

0

Methylene Chloride

Cohort Study—Mortality

Shannon et al., 1988

Lamp-manufacturing workers (primary exposure to methylene chloride), employed >6 months

1

1.47 (0.04–8.19)

Unspecified Mixtures of Organic Solvents

Cohort Study—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Mixed solvents, potential routine exposure

2

0.57 (0.07–2.07)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.24 Selected Epidemiologic Studies—Uterine and Endometrial Cancer and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene—Uterine

Cohort Study—Incidence

Hansen et al., 2001

Biologically monitored workers in Denmark

1

1.0 (0.01–5.4)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California, potential routine exposure

1

0.64 (0.02–3.57)

Morgan et al., 1988

Aerospace workers in Arizona, ever exposed

1

0.16 (0.00–0.91)

Tetrachloroethylene and Dry-cleaning Solvents—Uterine

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California, potential routine exposure

0

Blair et al., 1990

Dry-cleaning union members in Missouri

8

1.0 (0.4–2.0)

Methylene Chloride—Uterine/Endometrial

Cohort Studies—Mortality

Shannon et al., 1988

Lamp-manufacturing workers (primary exposure to methylene chloride), employed >6 months

2

2.14 (0.26–7.60)

Unspecified Mixtures of Organic Solvents—Uterine

Cohort Study—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Mixed solvents, potential routine exposure

1

0.31 (0.01–1.71)

UROLOGIC CANCERS

Description of Case-Control Studies

The characteristics of the case-control studies considered by the committee in drawing its conclusions of association are described below for each cancer site. The principal strengths and limitations of the studies are discussed below by cancer site.

Epidemiologic Studies of Exposure to Organic Solvents and Prostate Cancer

Gérin and co-workers (1998) evaluated 15 cancer risks, including the risk of prostate cancer, related to such occupational exposures as the hydrocarbons benzene, toluene, xylene, and styrene. The study had excellent information on exposures, as assessed by in-depth interviews that were coded blindly by a team of chemists and industrial hygienists, and sufficient information on most risk factors (see Table 6.25).

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.25 Description of Case-Control Study of Prostate Cancer and Exposure to Organic Solvents

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Gérin et al., 1998

Male cases, age 35–75 years, diagnosed in one of 19 large Montreal-area hospitals in 1979–1985 and histologically confirmed; controls identified concurrently at 18 other cancer sites; age-matched, population-based controls were also chosen from electoral lists and random-digit dialing

449

1066, consisting of 533 population controls and 533 randomly selected subjects from other cases of cancer

Benzene

Toluene

Xylene

In-person interviews (direct or proxy) with segments on work histories (job titles and self-reported exposures); analyzed and coded by a team of chemists and industrial hygienists (about 300 exposures on semiquantitative scales)

Logistic regression

Age, family income, ethnicity, cigarette smoking, respondent status

Response rates: 82% of all cases, 71% of population controls

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Prostate cancer mortality and incidence were not associated with exposure to trichloroethylene in the cohort studies conducted in the aircraft industry (Blair et al., 1998: SMR=1.0, 95% CI=0.5–2.1 for 5–25 unit-years of exposure; Boice et al., 1999: SMR=1.03, 95% CI=0.70–1.45). Morgan and colleagues (1998) found an excess risk of 18% (SMR=1.18, 95% CI=0.73–1.80), and the relative risks did not increase with increasing exposure to trichloroethylene. Other studies of occupationally exposed workers also found no association with exposure to trichloroethylene, including studies by Greenland and colleagues (1994) (OR=0.82, 95% CI=0.46–1.46), Hansen and colleagues (2001) (SIR=0.6, 95% CI=0.2–1.3), and Wilcosky and colleagues (1984) (OR=0.62 [CI not provided by the authors, and the committee was unable to calculate it]). A cohort study of US uranium-processing workers found no increase in prostate cancer when the subjects were exposed to “light” amounts of trichloroethylene with increasing years of exposure latency (SMR ranged from 0.78–1.04). After “moderate” exposure, risk increased; the SMR ranged from 1.35 (95% CI=0.17–10.4) after more than 2 years of exposure and no latency to 1.96 (95% CI=0.25–15.6) after more than 5 years of exposure and a 15-year latency (Ritz, 1999).

Axelson and co-workers (1994) reported an increased risk of prostate cancer among Swedish men occupationally exposed to trichloroethylene (RR=1.25, 95% CI=0.84–1.84). Anttila and colleagues (1995) reported an increased risk of prostate cancer after exposure to trichloroethylene (RR=1.38, 95% CI=0.73–2.35).

Two cohort studies of dry-cleaning workers did not show a positive association between tetrachloroethylene and risk of prostate cancer (Blair et al., 1990: SMR=0.7, 95% CI 0.2–1.7; Ruder et al., 1994: SMR=0.82, 95% CI 0.33–1.69).

A German study of rotogravure printers showed no association between prostate cancer and exposure to toluene (SMR=0.67, 95% CI=0.13–2.66) (Wiebelt and Becker, 1999). No increased risk of prostate cancer was found in workers in Montreal who reported exposure to toluene (ORhigh=0.4, 95% CI=0.1–1.4) (Gérin et al., 1998). A study of white, male rubber workers in Ohio found an association between exposure to toluene and prostate cancer (OR=2.6) (Wilcosky et al., 1984).

Both the Montreal study and the rubber-workers study evaluated prostate cancer risk and exposure to xylene: Gérin and co-workers (1998) reported an imprecise estimate of effect of “high” exposure to xylene (OR=1.4, 95% CI=0.5–4.0), and Wilcosky and colleagues (1984) reported an OR of 1.5.

Gérin and colleagues also examined exposure to benzene and risk of prostate cancer. An association was found with “medium” exposure to benzene (OR=1.7, 95% CI=0.9–3.0) but not “high” exposure (OR=0.9, 95% CI=0.4–2.1).

Only one cohort study of US cellulose-fiber production workers reported exposure to methylene chloride. An increased risk of prostate cancer was found (SMRhigh=1.79, 95% CI=0.95–3.06; SMRlow=1.40, 95% CI=0.64–2.66; SMRno exposure=1.04, 95% CI=0.22–3.05) (Gibbs et al., 1996). The magnitude of the risk increased in the high exposure group, with 20 years or more of latency (SMR=2.08; p<0.05) and with 20 years or more of exposure with 20 years of latency (SMR=2.91; p<0.05).

No increased risk of prostate cancer with exposure to unspecified mixtures of solvents was apparent in the studies the committee reviewed except one (Anttila et al., 1995). Studies that showed no increase include cohort studies by Boice and colleagues (1999) (SMR=1.0, 95% CI 0.78–1.26), Garabrant and colleagues (1988) (SMR=0.93,

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

95% CI=0.60–1.37), Matanoski and colleagues (1986) (SMR=0.99, 95% CI=0.82–1.18), Morgan and colleagues (1981) (SMR=0.84 [CI not provided by the study authors, and the committee was unable to calculate it]), and Greenland and colleagues (1994) (OR=0.84, 95% CI=0.49–1.42). The cohort study by Anttila and colleagues (1995) of Finnish workers monitored for exposure to halogenated hydrocarbons showed an SIR of 1.38 (95% CI=0.76–2.32).

Summary and Conclusion

Results of several large cohort studies of trichloroethylene exposed workers did not support an association between exposure and risk of prostate cancer, nor did the cohort studies of dry-cleaning workers. Although one positive study was identified for exposure to toluene, xylene, and benzene individually, other studies did not find an association, or the studies lacked any evidence of an exposure-response relationship. For exposure to methylene chloride, one study provided evidence for increased risk of prostate cancer with increasing years of exposure and latency, but other corroborating studies were not found. All but one of the studies on solvent mixtures found a positive association, therefore the committee was not able to determine whether an association exists between exposure and prostate cancer. Table 6.26 provides the key data points for each exposure reviewed by the committee in drawing its conclusion.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to solvents under review and prostate cancer.

TABLE 6.26 Selected Epidemiologic Studies—Prostate Cancer and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Studies—Incidence

Hansen et al., 2001

Biologically monitored Danish workers, ever exposed

6

0.6 (0.2–1.3)

Blair et al., 1998

Aircraft-maintenance workers in Utah

 

 

No exposure

61

1.0 (0.7–1.4)

 

<5 unit-years

64

1.1 (0.7–1.6)

 

5–25 unit-years

38

1.0 (0.6–1.6)

Anttila et al., 1995

Finnish workers monitored for exposure

 

 

Entire period since first measurement

13

1.38 (0.73–2.35)

 

0–9 years

2

1.09 (0.13–3.91)

 

10–19 years

3

0.56 (0.12–1.64)

 

20+ years

8

3.57 (1.54–7.02)

Axelson et al., 1994

Swedish men occupationally exposed, trichloroethylene

26

1.25 (0.84–1.84)

Cohort Studies—Mortality

Ritz, 1999

White male US uranium-processing workers

 

 

Duration of exposure, exposure lag

 

 

Trichloroethylene, light exposure

 

 

>2 years, no lag

10

0.78 (0.33–1.85)

 

>2 years, 15-year lag

10

0.91 (0.38–2.18)

 

>5 years, no lag

8

0.83 (0.33–2.09)

 

>5 years, 15-year lag

8

1.04 (0.40–2.70)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

 

Trichloroethylene, moderate exposure

 

 

>2 years, no lag

1

1.35 (0.17–10.4)

 

>2 years, 15-year lag

1

1.44 (0.19–11.4)

 

>5 years, no lag

1

1.58 (0.20–12.5)

 

>5 years, 15-year lag

1

1.96 (0.25–15.6)

Boice et al., 1999

Aircraft-manufacturing workers in California, routine exposure

32

1.03 (0.70–1.45)

Blair et al., 1998

Aircraft-maintenance workers in Utah

 

 

No trichloroethylene exposure

33

1.2 (0.7–2.1)

 

<5 unit-years

19

0.9 (0.5–1.8)

 

5–25 unit-years

13

1.0 (0.5–2.1)

Morgan et al., 1998

Aerospace workers in Arizona

 

 

Any exposure

21

1.18 (0.73–1.80)

 

Low exposure

7

1.29 (0.52–2.66)

 

High exposure

14

1.13 (0.62–1.89)

Greenland et al., 1994

White male US transformer manufacturers, ever exposed

NA

0.82 (0.46–1.46)

Wilcosky et al., 1984

White male rubber workers in Ohio, exposed >1 year

3

0.62

Tetrachloroethylene and Dry-Cleaning Solvents

Cohort Studies—Mortality

Ruder et al., 1994

Dry-cleaning labor-union workers

7

0.82 (0.33–1.69)

Blair et al., 1990

Members of a dry-cleaning union in St. Louis, MO

5

0.7 (0.2–1.7)

Toluene

Cohort Studies—Mortality

Wiebelt and Becker, 1999

Male German rotogravure printers, Employed >1 year

2

0.67 (0.13–2.66)

Wilcosky et al., 1984

White male rubber workers in Ohio, exposed >1 year

3

2.6

Case-Control Study

Gérin et al., 1998

Male residents of Montreal, Canada

 

 

Low exposure

51

1.0 (0.7–1.5)

 

Medium exposure

17

1.3 (0.7–2.5)

 

High exposure

3

0.4 (0.1–1.4)

Xylene

Cohort Study—Mortality

Wilcosky et al., 1984

White male rubber workers in Ohio, exposed >1 year

8

1.5

Case-Control Study

Gérin et al., 1998

Male residents of Montreal, Canada

 

 

Low exposure

46

1.2 (0.8–1.8)

 

Medium exposure

11

0.8 (0.4–1.7)

 

High exposure

6

1.4 (0.5–4.0)

Benzene

Cohort Study—Mortality

Wilcosky et al., 1984

White male rubber workers in Ohio, exposed >1 year

11

0.73

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Case-Control Study

Gérin et al., 1998

Male residents of Montreal, Canada

 

 

Low exposure

64

1.1 (0.8–1.5)

 

Medium exposure

22

1.7 (0.9–3.0)

 

High exposure

9

0.9 (0.4–2.1)

Methylene Chloride

Cohort Study—Mortality

Gibbs et al., 1996

Cellulose-fiber production workers

 

 

High exposure

13

1.79 (0.95–3.06)

 

≥20-year latency

NA

2.08 (p<0.05)

 

≥20-year latency and ≥20-year duration

NA

2.91 (p≤0.05)

 

Low exposure

9

1.40 (0.64–2.66)

 

No exposure

3

1.04 (0.22–3.05)

Unspecified Mixtures of Organic Solvents

Cohort Study—Incidence

Anttila et al., 1995

Finnish workers monitored for exposure

 

 

Halogenated hydrocarbons

14

1.38 (0.76–2.32)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Mixed solvents, routine exposure

70

1.0 (0.78–1.26)

 

Years exposed

 

 

<1

31

0.99 (0.65–1.49)

 

1–4

64

0.81 (0.59–1.13)

 

≥5

139

0.77 (0.58–1.02)

 

 

 

p-trend=0.06

Greenland et al., 1994

White male US transformer-assembly workers

 

Solvents, ever exposed

NA

0.84 (0.49–1.42)

Garabrant et al., 1988

Aircraft-manufacturing workers in California

25

0.93 (0.60–1.37)

Matanoski et al., 1986

US painters and allied tradesmen union members

117

0.99 (0.82–1.18)

Morgan et al., 1981

Male US paint and coatings manufacturers, employed >1 year

29

0.84

NOTE: NA=not available.

Epidemiologic Studies of Exposure to Organic Solvents and Bladder Cancer

All but one (Aschengrau et al., 1993) of the case-control studies of bladder cancer reviewed by the committee used occupational history to assess exposure, and in some studies information on specific chemical exposures was also obtained (Gérin et al., 1998; Pesch et al., 2000a). The study by Aschengrau and colleagues assessed exposure on the basis of estimated doses of tetrachloroethylene found in public drinking water in five towns of Cape Cod, Massachusetts. Two studies included interviews with proxies if subjects were too ill to be interviewed (Morrison et al., 1985; Teschke et al., 1997). Most studies simply grouped exposure defined broadly on the basis of occupation, including work in the painting industry (Cordier et al., 1993; Jensen et al., 1987; La Vecchia et al., 1990; Morrison et al., 1985; Vineis and Magnani, 1985), in laundry and dry-cleaning services (Silverman et al., 1989a,b; Smith et al., 1985), and in both fields (Schoenberg et al., 1984; Teschke et al.,

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

1997). Very low response rates were found in the study by Risch and colleagues (1988) of exposure to paints. A number of case-control studies used self-reported information on exposures that were not otherwise validated (Jensen et al., 1987; La Vecchia et al., 1990; Risch et al., 1988; Schoenberg et al., 1984; Smith et al., 1985; Vineis and Magnani, 1985) and thus may be subject to recall bias if the controls were not ill.

The main accepted risk factor for bladder cancer is cigarette smoking, and all but one (Smith et al., 1985) of the case-control studies of bladder cancer and exposure to specific and unspecified organic solvents reviewed by the committee accounted for smoking in some way. In most studies, subjects were asked whether they currently smoked; others inquired about the number of cigarettes smoked per day (Aschengrau et al., 1993), lifetime cigarette consumption (Risch et al., 1988), or duration of cigarette smoking (Schoenberg et al., 1984).

Case-control studies of bladder cancer and exposure to organic solvents that had reasonably good assessments of exposure, adequate control for confounding, and histologic confirmation of outcome include those by: Aschengrau et al., 1993; Cordier et al., 1993; Gérin et al., 1998; Pesch et al., 2000a; and Silverman et al., 1989a,b (see Table 6.27).

In addition to the case-control study by Pesch and colleagues (2000a) described above, there were several cohort studies of biologically monitored workers, transformer manufacturers, and aircraft and aerospace workers. One study of US aircraft-maintenance workers showed an RRany exposure of 1.2 (95% CI=0.5–2.9) for bladder cancer (Blair et al., 1998); another showed no association (RR=0.55, 95% CI=0.18–1.28) (Boice et al., 1999). An increased risk of bladder cancer also was observed in a cohort of US aerospace-manufacturing workers (SMR=1.36, 95% CI=0.59–2.68) (Morgan et al., 1998). Three other cohort studies of workers biologically monitored for exposure to trichloroethylene, as shown by the presence of a urinary metabolite, showed no association (relative risks ranged from 0.61 to 1.1) between this marker and bladder cancer (Anttila et al., 1995; Axelson et al., 1994; Hansen et al., 2001). A nested case-control study within a cohort of transformer-assembly facility workers did not find an association between risk of bladder cancer and exposure to trichloroethylene (OR=0.85, 95% CI=0.32–3.32) (Greenland et al., 1994).

The case-control study by Pesch and colleagues (2000a) suggested a positive association with urothelial carcinoma (a cancer of the urinary tract that affects mostly the bladder). Using a job-exposure matrix, the authors found a 10% excess risk (OR=1.1; 95% CI=0.9–1.4) among men with “high” exposure and a 60% excess risk among women with “high” exposure (OR=1.6, 95% CI=1.0–2.5). On the basis of the job task-exposure matrix, the magnitude of the OR for men with “high” exposure was increased to 1.3 (95% CI=0.9–1.7). The job task-exposure matrix was not used to evaluate exposure of women.

The committee used two cohort studies and several case-control studies of dry-cleaning workers and other workers exposed to tetrachloroethylene in determining whether there is an association between exposure and bladder cancer. The two dry-cleaner cohort studies showed increased mortality from bladder cancer: Blair and colleagues (1990) reported an increased SMR of 1.7 (95% CI=0.7–3.3), and Ruder and colleagues (2001) found a 122% increase in mortality from bladder cancer (SMR=2.22, 95% CI=1.06–4.08). The magnitude of the association was higher (SMR=4.31, 95% CI=1.85–8.76) among workers employed more than 5 years and in whom 20 years had passed since the first exposure.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.27 Description of Case-Control Studies of Bladder Cancer and Exposure to Organic Solvents

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Schoenberg et al., 1984

Male cases, age 21–84 years, with histologically confirmed diagnosis in New Jersey in February 1978–1979; controls identified through RDD (age 21–64 years) and HCFA records (age 65–85 years), stratified for age

658

1258

Painting or artistic work

Dry-cleaning work

In-person interviews with questionnaires assessing lifetime occupational history (job titles)

Logistic regression

Age, duration of cigarette smoking, other occupations

Response rates: 89.7% of cases, 86.6% of controls

Morrison et al., 1985

Cases, age 21–89 years, identified at hospitals in Boston; Greater Manchester County, UK; and Nagoya, Japan, in 1976–1978; controls selected from respective areas’ electoral registers, matched for age and sex

430 Boston

399 UK

226 Japan

397 Boston

493 UK

443 Japan

Painting work

In-person interview (direct or proxy) assessing occupational history (job titles); job titles were coded

Logistic regression

Age, cigarette smoking

No response rates provided

Smith et al., 1985

Cases and controls, age 21–84 years, residing in the nine SEER population-based areas and New Jersey, who participated in the NCI National Bladder Cancer Study; cases with histologically confirmed cancer; controls frequency matched for age and sex

7748 total participants with and without bladder cancer, classified as: worked in dry-cleaning operations (N=103), experienced related exposures (N=5776), or neither (unexposed; N=1869)

Laundry and dry-cleaning work

In-person interview with structured questionnaire regarding occupational history (job or industry titles)

Logistic regression

Age, sex

No response rates provided

Vineis and Magnani, 1985

Cases, age less than 70 years, identified from the Main Hospital in Torino, Italy, in 1978–1983; controls from same hospital diagnosed with benign urologic conditions, matched for age

512

596

Painting work

In-person interviews regarding lifetime occupational history (job or industry titles) obtained in hospital and blindly coded

Mantel-Haenszel

Age, smoking

No response rates provided

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Jensen et al., 1987

Cases reported to the Danish Cancer Registry from Copenhagen area in 1979–1981, with 99% histologic verification; controls selected from the residents of the municipalities from which cases arose, matched for sex and age

371

771

Painting work

In-person interviews with questionnaire assessing occupational history (job titles and self-reported exposures)

Logistic regression

Age, sex, smoking

Response rates: 94.4% of cases, 75.1% of controls

Risch et al., 1988

Cases, age 35–79 years, identified through a combination of cancer registry reporting and hospital record review in four cities in Canada in 1979–1982 with histologic confirmation; controls selected randomly from population listings and matched on birth year, sex, and residence area

826

792

Organic solvents

Paints

In-person interview with questionnaire assessing specific occupational exposures (self-reports)

Conditional logistic regression

Matching variables, lifetime cigarette consumption

Response rates: 67% of cases, 53% of controls

Silverman et al., 1989a

White cases, age 21–84 years, in 10 US areas in 1977–1978 with histologic confirmation; controls identified through RDD (age 21–64 years) and HCFA records (age 65–84 years), matched for age and geographic area

2100

3874

Dry-cleaning, ironing, and pressing work

Questionnaire administered by in-person interview (job or industry titles); industries and job titles coded by study authors and grouped by potential exposures

Logistic regression

Smoking, age

No response rates provided

Silverman et al., 1989b

Nonwhite cases, same study as above (Silverman et al., 1989a)

126

383

See above

See above

See above

See above

La Vecchia et al., 1990

Cases, age less than 75 years, admitted to NCI or clinics and hospitals in Milan, Italy, in 1985–1988 with histologic confirmation; controls admitted to the same network of hospitals for acute nonneoplastic conditions

263

287

Painting work

Chemical-industry work

Structured questionnaire to assess lifetime employment in 19 industries or occupations and 14 specific agents (job or industry titles and self-reported exposures)

Mantel-Haenszel

Age, sex, smoking

Response rate: >97% of cases and controls

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Aschengrau et al., 1993

Cases reported to the Massachusetts Cancer Registry, diagnosed in 1983–1986 among residents of five upper Cape Cod towns; living controls were selected from HCFA records and through RDD; deceased controls identified by the state Department of Vital Statistics and Research files

61

852

Tetrachlorid e-ethylene

Exposure dose estimated in areas of contaminated drinking water, accounting for location and years of residence, water flow, pipe characteristics

Logistic regression

Sex, age at diagnosis, vital status, educational level, usual number of cigarettes smoked, occupational exposure to solvents, specific cancer risk factors controlled for in respective analyses

Response rates: 80.6% of cases, 75.9% of HCFA controls, 73.9% of RDD controls, 78.8% of next of kin of deceased controls

Cordier et al., 1993

Cases, under age 80 years, from seven French hospitals in 1984–1987, with histologic confirmation; controls selected from the same hospitals from patients admitted for causes other than cancer, respiratory disease, or symptoms related to bladder cancer, matched for sex, age, ethnicity, and residence

765

765

Solvents

Painting work

In-person interviews with segments on work histories (job titles); analyzed and coded by a team of experts in industrial hygiene

Logistic regression

Smoking status, hospital, age, place of residence

No response rates provided

Teschke et al., 1997

Cases, age 19 years and over, registered with the British Columbia Cancer Agency in 1990–1991 with histologic confirmation; controls selected from the provincial voters list and matched on age and sex

105

139

Laundry personnel

Painters

Occupational histories (job titles) and self-reported exposures obtained (direct or proxy) through standardized questionnaire (in-person or telephone interview)

Adjusted ORs

Sex, age, cigarette smoking

Response rates: 88.2% of cases, 80.3% of controls

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Gérin et al., 1998

Male cases, age 35–75 years, diagnosed in one of 19 large Montreal-area hospitals in 1979–1985 and histologically confirmed; controls identified concurrently at 18 other cancer sites; age-matched, population-based controls were also chosen from electoral lists and random-digit dialing

484

1066, consisting of 533 population controls and 533 randomly selected subjects from other cases of cancer

Benzene

Toluene

Xylene

In-person interviews (direct or proxy) with segments on work histories (job titles and self-reported exposures); analyzed and coded by a team of chemists and industrial hygienists (about 300 exposures on semiquantitative scales)

Logistic regression

Age, family income, ethnicity, cigarette smoking, respondent status

Response rates: 82% of all cases, 71% of population controls

Pesch et al., 2000a

Cases from large hospitals in five regions in Germany in 1991–1995 with histologic confirmation; controls randomly selected from local residency registries matched on region, sex, and age

1035 urothelial

4298

Trichloroethylene

Ethylene-ethylene

Benzene

Organic solvents

In-person interviews of lifetime occupational history using questionnaire to assess job titles and self-reported exposures; exposures ascertained by job-exposure matrixes

Logistic regression

Age, study center, smoking

Response rates: 84% of cases, 71% of controls

NOTE: HCFA=Health Care Financing Administration; NCI=National Cancer Institute; SEER=Surveillance, Epidemiology, and End Results; RDD=random-digit dialing.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Several case-control studies reported increased relative risks of bladder cancer with exposure in the dry-cleaning industry. Schoenberg and colleagues (1984) showed an OR of 1.33 (95% CI=0.50–3.58), and Teschke and colleagues (1997) found an OR of 2.3 (95% CI=0.4–13.9) in people ever employed as laundry personnel. Smith and co-workers (1985) showed an increased risk of bladder cancer among laundry and dry-cleaning workers, and this risk was higher among exposed smokers (OR=3.94, 95% CI=2.39–6.51) than among exposed nonsmokers (OR=1.31, 95% CI=0.85–2.03).

On the basis of exposure to tetrachloroethylene in public drinking water, Aschengrau and colleagues (1993) showed an OR of 1.16 (95% CI=0.48–2.48) for low levels of exposure and an OR of 6.04 (95% CI=1.32–21.84) for high levels of exposure. Using a job-exposure matrix, Pesch and colleagues (2000a) demonstrated an increased risk in men exposed to high levels of tetrachloroethylene (OR=1.2, 95% CI=1.0–1.5) and no increased risk in women (OR=1.0, 95%CI=0.6–1.9).

Two studies provided data on exposure to benzene and bladder cancer. Exposure to benzene, as assigned by industrial hygienists after reviewing detailed job histories, was not associated with bladder cancer in the Montreal case-control study (Gérin et al., 1998). In the study by Pesch and colleagues (2000a), there was an indication among men of risks increasing with increasing exposure on the basis of three methods of assessing exposure. For example, for “substantial” exposure to benzene, according to a British job-exposure matrix (Pannett et al., 1985), the RR of bladder cancer in men was 1.5 (95% CI=1.0–2.1) and in women 1.4 (95% CI =0.6–3.3) (Pesch et al., 2000a). However, the risks did not increase with increasing exposure among women according to the other two exposure-determination methods.

Three small occupational cohort studies (Svensson et al., 1990; Walker et al., 1993; Wiebelt and Becker, 1999) and a population-based case-control study (Gérin et al., 1998) found no evidence of an association between bladder cancer and exposure to toluene or xylene. Most of the estimates of RR were below unity, and the highest was 1.20 (95% CI=0.25–3.51) in women working as shoe manufacturers (Walker et al., 1993).

The study that had the largest number of exposed cases was the cohort study of US painters and other union members that showed an increased risk of bladder cancer among painters (SMR=1.23, 95% CI=1.05–1.43) but not in nonpainters (SMR=0.74, 95% CI=0.46–1.11) (Steenland and Palu, 1999). A formal Poisson regression analysis comparing painters and nonpainters showed an RR of 1.77 (95% CI=1.13–2.77) in painters. A smaller study of US painters (Matanoski et al., 1986) and a small study of paint-manufacturing workers (Morgan et al., 1981) showed no increased risk of bladder cancer (RR=1.06, 95% CI=0.78–1.41 and RR=0.98 [no CI was available, and the committee was not able to calculate it], respectively). Boice and colleagues (1999) did not find an association in aircraft-manufacturing workers (SMR=0.85, 95% CI=0.49–1.35). Anttila and colleagues (1995) also found no increased risk of bladder cancer (SIR=0.73, 95% CI=0.24–1.71) among Finnish workers occupationally exposed to halogenated hydrocarbons. A cohort study of US aircraft-manufacturing workers in California showed a 26% excess risk of bladder cancer (SMR=1.26, 95% CI=0.74–2.03) (Garabrant et al., 1988). The study of US transformer-assembly workers showed a 21% increase in mortality (OR=1.21, 95% CI=0.49–2.98) (Greenland et al., 1994).

Pesch and colleagues (2000a) found an increased risk in painters in Germany who had a “very long” duration of exposure (OR=1.6, 95% CI=0.5–4.7) and in men who had “substantial” exposure to paints (OR=1.6, 95% CI=1.1–2.3). The authors did not show that the risk in painters increased with increasing duration of exposure (medium: OR=1.3, 95% CI=

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

0.6–2.6; long: OR=0.7, 95% CI=0.3–1.61; very long: OR=1.6, 95% CI=0.5–4.7). Those analyses were adjusted for smoking. In British Columbia (Teschke et al., 1997), the risk was increased in people who had ever been employed as painters (OR=2.8, 95% CI=0.4–21.3) and after 20 years since first employment (OR=2.0, 95% CI 0.1–33.0). The risk of bladder cancer increased in the Copenhagen study (OR=2.54, 95% CI=1.12–5.73) and increased with increasing duration of employment (1–19 years: OR=1.6, 95% CI=0.5–5.5; 20 years or more: OR=4.1, 95% CI=1.2–13.9) (Jensen et al., 1987). Positive associations were also found in studies in Milan (OR=1.8, 95% CI=0.8–3.7) (La Vecchia et al., 1990), Canada (OR=1.18, 95% CI=0.87–1.62) (Risch et al., 1988), Boston (OR=1.5, 95% CI 0.9–2.4) (Morrison et al., 1985), and New Jersey (OR=1.53, 95% CI=0.96–2.44) (Schoenberg et al., 1984).

Summary and Conclusions

Several studies of trichloroethylene and bladder cancer showed weak and imprecise associations. Most suffered from low statistical power and probable exposure misclassification, so the committee concluded that there was insufficient evidence to determine whether an association exists.

For exposure to tetrachloroethylene and dry-cleaning solvents, a number of cohort and case-control studies showed a positive association between exposure and risk of bladder cancer. Therefore, the committee judged that the data, though limited, was suggestive of an association.

The committee concludes, from its assessment of the epidemiologic literature, that there is limited/suggestive evidence of an association between chronic exposure to tetrachloroethylene and dry-cleaning solvents and bladder cancer.

Based on the consistency of association in the case-control studies (Jensen et al., 1987; Morrison et al., 1985; Pesch et al., 2000b; Schoenberg et al., 1984) and the positive findings in the US cohorts of painters (Steenland and Palu, 1999) and aircraft workers (Garabrant et al., 1988), the committee decided that the evidence between exposure to unspecified mixtures of organic solvents and bladder cancer was both limited and suggestive of an association.

The committee concludes, from its assessment of the epidemiologic literature, that there is limited/suggestive evidence of an association between chronic exposure to unspecified mixtures of organic solvents and bladder cancer.

In contrast, the committee considered the findings for exposure to benzene and risk of bladder cancer to be inconsistent. The studies on toluene and xylene did not find an association. Table 6.28 identifies the studies used by the committee in making its conclusions, and unless indicated in the table, the study populations include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to other solvents under review and bladder cancer.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.28 Selected Epidemiologic Studies—Bladder Cancer and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Studies—Incidence

Hansen et al., 2001

Biologically monitored Danish workers

 

 

Males, ever exposed

10

1.1 (0.5–2.0)

 

Females, ever exposed

0

Blair et al., 1998

Aircraft maintenance in Utah

 

 

No trichloroethylene exposure

10

1.3 (0.5–3.5)

 

<5 unit-years

13

1.7 (0.6–4.4)

 

5–25 unit-years

9

1.7 (0.6–4.9)

Anttila et al., 1995

Finnish workers occupationally exposed

 

 

Entire period since first measurement

5

0.82 (0.27–1.90)

 

0–9

1

0.65 (0.02–3.59)

 

10–19

2

0.61 (0.07–2.22)

 

20+

2

1.51 (0.18–5.44)

Axelson et al., 1994

Swedish men, occupationally exposed

8

1.02 (0.44–2.00)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California, routine exposure

5

0.55 (0.18–1.28)

Blair et al., 1998

Aircraft maintenance workers in Utah

 

 

Any exposure

17

1.2 (0.5–2.9)

 

No trichloroethylene exposure

4

0.7 (0.2–2.8)

 

<5 unit-years

7

1.8 (0.5–6.2)

 

5–25 unit-years

5

2.1 (0.6–8.0)

 

Low-level intermittent

10

1.5 (0.4–4.8)

 

Low-level continuous

9

2.0 (0.6–6.4)

Morgan et al., 1998

Aerospace workers in Arizona

 

Any exposure

8

1.36 (0.59–2.68)

 

Low exposure

1

0.51 (0.01–2.83)

 

High exposure

7

1.79 (0.72–3.69)

Greenland et al., 1994

White male US transformer manufacturers, ever exposed

NA

0.85 (0.32–3.32)

Case-Control Study

Pesch et al., 2000a

Participants in multiple centers in Germany

 

 

German job-exposure matrix

 

 

Trichloroethylene (males)

 

 

Medium

154

1.1 (0.8–1.3)

 

High

182

1.1 (0.9–1.4)

 

Substantial

68

1.3 (0.9–1.7)

 

Trichloroethylene (females)

 

 

Medium

21

1.0 (0.6–1.7)

 

High

32

1.6 (1.0–2.5)

 

Substantial

3

0.6 (0.2–2.3)

 

Job task-exposure matrix approach

 

 

Trichloroethylene (males)

 

 

Medium

47

0.8 (0.6–1.2)

 

High

74

1.3 (0.9–1.7)

 

Substantial

36

1.8 (1.2–2.7)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Tetrachloroethylene and Dry-cleaning Solvents

Cohort Studies—Mortality

Ruder et al., 2001

Dry-cleaning labor-union workers

10

2.22 (1.06–4.08)

 

≥5 years exposure with >20 years latency

8

4.31 (1.85–8.76)

 

Tetrachloroethylene plus other solvents (likely Stoddard)

10

3.15 (1.51–5.79)

Blair et al., 1990

Members of a dry-cleaning union in St. Louis, MO

8

1.7 (0.7–3.3)

Case-Control Studies

Pesch et al., 2000a

Participants in multiple centers in Germany

 

 

German job-exposure matrix

 

 

Tetrachloroethylene (males)

 

 

Medium

162

1.1 (0.9–1.3)

 

High

172

1.2 (1.0–1.5)

 

Substantial

71

1.4 (1.0–1.9)

 

Tetrachloroethylene (females)

 

 

Medium

21

1.8 (1.0–3.0)

 

High

16

1.0 (0.6–1.9)

 

Substantial

3

0.7 (0.2–2.5)

 

Job task-exposure matrix approach

 

 

Tetrachloroethylene (males)

 

 

Medium

37

1.0 (0.7–1.5)

 

High

47

1.2 (0.8–1.7)

 

Substantial

22

1.8 (1.1–3.1)

Teschke et al., 1997

Residents of British Columbia, Canada

 

Laundry personnel, ever employed

5

2.3 (0.4–13.9)

 

Laundry personnel, most recent 20 years removed

4

1.8 (0.3–11.3)

Aschengrau et al., 1993

Residents of upper Cape Cod, MA

 

Any exposure (no latency)

13

1.55 (0.74–3.01)

 

Low exposure

9

1.16 (0.48–2.48)

 

High exposure

4

6.04 (1.32–21.84)

Silverman et al., 1989b

Nonwhite males in 10 US areas

 

Dry cleaner, ironer, or presser, ever employed

11

2.8 (1.1–7.4)

 

<5 years duration

7

5.3

 

5+ years duration

4

1.8

Smith et al., 1985

Incident cases in 10 US areas

 

 

Laundry and dry cleaners, ever employed

 

 

Nonsmoker

NA

1.31 (0.85–2.03)

 

Former smoker

NA

2.99 (1.80–4.97)

 

Current smoker

NA

3.94 (2.39–6.51)

 

Chemically related exposure group

 

 

Nonsmoker

NA

1.11 (0.99–1.25)

 

Former smoker

NA

2.01 (1.69–2.40)

 

Current smoker

NA

3.12 (2.62–3.71)

Schoenberg et al., 1984

Male residents of New Jersey

 

Dry-cleaning, ever employed

7

1.33 (0.50–3.58)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Benzene

Case-Control Studies

Pesch et al., 2000a

Participants in multiple centers in Germany

 

 

 

British job-exposure matrix

 

 

 

Benzene (males)

 

 

 

Medium

95

1.1 (0.8–1.4)

 

High

70

0.9 (0.7–1.2)

 

Substantial

47

1.5 (1.0–2.1)

 

Benzene (females)

 

 

 

Medium

21

1.2 (0.7–2.0)

 

High

18

1.5 (0.9–2.8)

 

Substantial

8

1.4 (0.6–3.3)

 

German job-exposure matrix

 

 

 

Benzene (males)

 

 

 

Medium

177

1.1 (0.9–1.3)

 

High

169

1.2 (1.0–1.6)

 

Substantial

68

1.2 (0.8–1.6)

 

Benzene (females)

 

 

 

Medium

27

1.0 (0.7–1.7)

 

High

23

0.8 (0.5–1.4)

 

Substantial

5

0.6 (0.2–1.6)

 

Job task-exposure matrix approach

 

 

 

Benzene (males)

 

 

 

Medium

51

0.7 (0.5–1.0)

 

High

71

1.0 (0.7–1.3)

 

Substantial

37

1.4 (0.9–2.1)

 

Benzene (females)

 

 

 

Medium

2

0.4 (0.1–1.8)

 

High

3

0.4 (0.1–1.2)

 

Substantial

2

0.8 (0.2–3.7)

Gérin et al., 1998

Male residents of Montreal, Canada

 

 

 

Low exposure

65

1.0 (0.7–1.3)

 

Medium exposure

22

1.2 (0.7–2.0)

 

High exposure

2

0.2 (0.0–0.6)

Toluene

Cohort Study—Incidence

Svensson et al., 1990

Male rotogravure printers in Sweden, employed >3 months

 

 

 

Bladder and kidney cancer

4

0.64 (0.18–1.65)

 

≥5 yrs exoposed with >10 yrs latency

4

0.85 (0.23–2.16)

Cohort Study—Mortality

Wiebelt and Becker, 1999

Male German rotogravure printers, employed >1 year

2

0.66 (0.08–3.27)

Walker et al., 1993

Shoe manufacturers in two plants in Ohio, potentially exposed to toluene and other solvents

 

 

 

Total

7

0.99 (0.40–2.05)

 

Males

4

0.87 (0.24–2.25)

 

Females

3

1.20 (0.25–3.51)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Svensson et al., 1990

Male rotogravure printers in Sweden, employed >3 months

 

 

 

Bladder and kidney cancer

1

0.45 (0.01–2.53)

 

≥5 yrs exoposed with >10 yrs latency

1

0.57 (0.01–3.20)

Case-Control Study

Gérin et al., 1998

Male residents of Montreal, Canada

 

 

Low exposure

52

0.9 (0.6–1.3)

 

Medium exposure

6

0.3 (0.1–0.7)

 

High exposure

7

1.0 (0.4–2.5)

Xylene

Case-Control Study

Gérin et al., 1998

Male residents of Montreal, Canada

 

 

Low exposure

35

0.8 (0.5–1.1)

 

Medium exposure

16

1.0 (0.5–1.8)

 

High exposure

3

0.8 (0.2–3.2)

Unspecified Mixtures of Organic Solvents

Cohort Study—Incidence

Anttila et al., 1995

Finnish workers monitored for exposure

 

 

Halogenated hydrocarbons

5

0.73 (0.24–1.71)

Cohort Studies—Mortality

Steenland and Palu, 1999

US painters and other union members

 

Painters

166

1.23 (1.05–1.43)

 

20 years since membership

146

1.25 (1.06–1.47)

 

Nonpainters

22

0.74 (0.46–1.11)

 

20 years since membership

19

0.84 (0.51–1.31)

 

Poisson regression comparing painters with nonpainters:

 

 

Rate ratio

166

1.77 (1.13–2.77)

 

Rate ratio, ≥20 years since membership

146

1.55 (0.96–2.51)

Greenland et al., 1994

White male US transformer-assembly workers

NA

1.21 (0.49–2.98)

 

Solvents, ever exposed

 

Garabrant et al., 1988

Aircraft-manufacturing workers in California

 

 

Entire cohort

17

1.26 (0.74–2.03)

 

25–29 years of employment

2

2.66 (0.32–9.50)a

 

30+ years of employment

1

2.45 (0.06–13.59)a

Matanoski et al., 1986

US painters and allied tradesmen union members

48

1.06 (0.78–1.41)

Morgan et al., 1981

Male US paint and coatings manufacturers, employed >1 year

16

0.98

Walker et al., 1993

Shoe-manufacturing workers in Ohio, potentially exposed

7

0.99 (0.40–2.05)

 

Males

4

0.87 (0.24–2.25)

 

Females

3

1.20 (0.25–3.51)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Case-Control Studies

Pesch et al., 2000a

Residents of multiple centers in Germany

 

 

Duration of exposure—painters:

 

 

Medium

12

1.3 (0.6–2.6)

 

Long

6

0.7 (0.3–1.6)

 

Very long

5

1.6 (0.5–4.7)

 

Job-exposure approach, organic solvents:

 

 

Males, substantial exposure

43

1.4 (0.9–2.0)

 

Females, substantial exposure

8

1.5 (0.6–3.5)

 

Job-exposure approach, paints:

 

 

Males, substantial exposure

38

1.6 (1.1–2.3)

 

Females, substantial exposure

2

0.3 (0.1–1.13)

 

Job task-exposure matrix, paints:

 

 

Males, substantial exposure

33

1.5 (1.0–2.3)

 

Females, substantial exposure

10

2.1 (1.0–4.4)

Teschke et al., 1997

Painters in British Columbia, Canada

 

Ever employed

4

2.8 (0.4–21.3)

 

20 years since employment

2

2.0 (0.1–33.0)

Cordier et al., 1993

Residents of France

 

 

Painter, ever employed

19

0.97 (0.50–1.88)

 

Solvents, ever exposed

171

1.28 (0.98–1.68)

La Vecchia et al., 1990

Residents of Milan, Italy

 

Painting, ever employed

NA

1.8 (0.8–3.7)

 

Chemical industry, ever exposed

NA

1.7 (0.9–3.3)

Silverman et al., 1989a

White males employed as painters

 

<5 years

50

1.7

 

5–9 years

14

0.9

 

10–24 years

26

1.6

 

25+ years

22

1.9

Risch et al., 1988

Residents of Canada

 

 

Organic solvents, ever exposed

208

1.14 (0.82–1.57)

 

Organic solvents, exposed 8–28 years

NA

1.03 (0.70–1.52)

 

Paints, ever exposed

204

1.18 (0.87–1.62)

 

Paints, exposed 8–28 years

NA

1.11 (0.77–1.60)

Jensen et al., 1987

Residents of Copenhagen, Denmark

 

 

Painting, ever employed

13

2.54 (1.12–5.73)

 

1–19 years

5

1.6 (0.5–5.5)

 

≥20 years

8

4.1 (1.2–13.9)

Vineis and Magnani, 1985

Painters in Turin, Italy

 

Painter in building industry

12

1.0 (0.4–2.2)

 

Painter in carpentry

1

0.6 (0.04–8.4)

 

Car painter

7

2.0 (0.6–7.0)

 

Spray-painter

2

1.2 (0.2–5.8)

Morrison et al., 1985

Residents occupationally exposed to paints

 

Boston, MA

35

1.5 (0.9–2.4)b

 

Manchester, UK

23

0.7 (0.5–1.2)b

 

Nagoya, Japan

5

0.7 (0.3–1.7)b

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Schoenberg et al., 1984

Male painters or artists in New Jersey, ever employed

39

1.53 (0.96–2.44)

NOTE: NA=not available.

a95% CI calculated by the committee with standard methods from the observed and expected numbers presented in the original study.

b90% CI.

Epidemiologic Studies of Exposure to Organic Solvents and Kidney Cancer

Exposure to organic solvents was determined by using responses from in-person interviews or structured questionnaires about job titles and occupational and industrial exposures. Many of the studies reviewed on kidney cancer included small numbers of cases, which limited their statistical power. Several of the studies with large numbers of cases used self-reports to determine exposure, including those by: Asal et al., 1988; Dosemeci et al., 1999; Jensen et al., 1988; Mandel et al., 1995; McCredie and Stewart, 1993; Mellemgaard et al., 1994; Schlehofer et al., 1995; Sharpe et al., 1989; and Vamvakas et al., 1998. Such exposure determinations may be subject to misclassification errors that would affect risk estimates.

Although several studies included assessment of risk associated with specific solvent exposure, exposure was based on job titles that were then linked to potential exposure information determined by a job-exposure matrix or by industrial hygienists (Harrington et al., 1989; Partanen et al., 1991; Pesch et al., 2000b). Risk factors for kidney cancer that would require consideration in analyses are not well established.

Like the studies of bladder cancer, the case-control studies considered by the committee to be of relatively high quality had good information on exposure, controlled adequately for confounding, and used histologic confirmation of outcomes (Aschengrau et al., 1993; Asal et al., 1988; Gérin et al., 1998; Partanen et al., 1991; and Pesch et al., 2000b) (see Table 6.29).

Kidney cancer was not associated with exposure to trichloroethylene in three studies of biologically monitored workers in Scandinavian countries (Anttila et al., 1995; Axelson et al., 1994; Hansen et al., 2001) or in a study of US transformer manufacturers (Greenland et al., 1994). In the study by Anttila and colleagues, the SIR was 0.87 (95% CI=0.32–1.89) in all workers exposed, 1.16 (95% CI=0.42–2.52) in Swedish men in Axelson and colleagues’ study, and 0.9 (95% CI=0.2–2.6) in Danish workers in Hansen and colleagues’ study. Greenland and colleagues showed an OR of 0.99 (95% CI=0.30–3.32) in the transformer cohort. Three cohort studies of workers in aircraft and aerospace industries were inconsistent; no association (SMR=0.99, 95% CI=0.40–2.04) was found in a California aircraft-manufacturing study (Boice et al., 1999). Studies of aircraft-maintenance workers in Utah (Blair et al., 1998) and aerospace workers in Arizona (Morgan et al., 1998) showed increased relative risks but no exposure-response relationships.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.29 Description of Case-Control Studies of Kidney Cancer and Exposure to Organic Solvents

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Asal et al., 1988

Cases from 29 hospitals in Oklahoma diagnosed and confirmed in 1981–1984; hospital controls selected from the same hospitals and matched on age, sex, race, hospital, and date of admission; population-based controls selected through RDD

315 renal cell carcinoma

313 hospital

336 population

Dry-cleaning work

Painter or paint-manufacturing work

In-person interview assessing occupations (job titles) and industrial exposures

Logistic regression

Weight, age, alcohol consumption, occupations, smoking, snuff use, coffee consumption, kidney stones, hypertension, other medical factors

 

No response rates provided

 

Jensen et al., 1988

Cases, under age 80 years, reported to the Danish Cancer Registry from Copenhagen and the surrounding island of Sjaelland in 1979–1982, with 90% histologic verification; controls selected from the hospitals from which cases arose, excluding those with urinary tract and smoking-related diseases; controls matched for hospital, sex, and age

96 renal pelvis and ureter

288

Painter or paint-manufacturing work

In-person interviews with questionnaire assessing personal habits and occupational history (job or industry titles and self-reported exposures)

Logistic regression

Sex, lifetime tobacco smoking

 

Response rates: 99.0% of cases, 100.0% of controls

 

Harrington et al., 1989

Cases diagnosed and histologically confirmed in 1984–1985 and reported to the West Midlands Regional Cancer Registry (UK); controls randomly selected from practitioner records and matched for age, sex, ethnicity, location, and socioeconomic group

54 renal

54

Solvents

In-person interviews with questionnaire assessing lifetime occupational history (job titles); exposure indexes determined by occupational hygienist or chemist

Matched analyses

None

 

No response rates provided

 

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Sharpe et al., 1989

Cases diagnosed at one of four Montreal-area hospitals in 1982–1986 and one of five other hospitals in 1982–1987; cases were histologically confirmed and alive at time of chart review; controls selected from suspected renal-cell carcinoma cases, but final diagnoses were not cancer; matched 1:1 for sex, age, and urologist

164 renal

161

Organic solvents

History of exposure to hydrocarbons obtained through mailed questionnaire and supplemented by telephone interview (self-reports)

Univariat e analysis

None

 

Response rate: 97% overall

 

Partanen et al., 1991

Cases, age over 20 years, identified through the Finnish Cancer Registry in 1977–1978; controls randomly selected from the Population Register Centre matched for year of birth, sex, and survival status

408 renal cell

819

Nonchlorinat ed solvents

Mailed questionnaire or phone interview (direct or proxy) assessing lifetime occupational history (job or industry titles); industrial hygienist coded and assigned summary indicators of specific exposures

Condition al logistic regression

Matching variables, smoking, coffee consumption, obesity

 

Response rates: 69% of cases, 68% of controls

 

Aschengrau et al., 1993

Cases reported to the Massachusetts Cancer Registry, diagnosed in 1983–1986 among residents of five upper Cape Cod towns; living controls were selected from HCFA records and RDD; deceased controls identified by the state Department of Vital Statistics and Research files

35 kidney

777

Ethylene-ethylene

Exposure dose estimated in areas of contaminated drinking water, accounting for location and years of residence, water flow, pipe characteristics

Logistic regression

Sex, age at diagnosis, vital status, educational level, usual number of cigarettes smoked, occupational exposure to solvents, specific cancer risk factors controlled for in respective analyses

 

Response rates: 80.6% of cases, 75.9% of HCFA controls, 73.9% of RDD controls, 78.8% of next of kin of deceased controls

 

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

McCredie and Stewart, 1993

Cases, age 20–79 years, among residents of New South Wales in 1989–1990 identified from the New South Wales Central Cancer Registry and from physicians; controls selected from electoral rolls and matched on age distribution

489 renal cell

147 renal pelvic

523

Dry-cleaning industry work

Questionnaire (in-person interview or mailed with telephone followup) to assess employment in specific occupations and industries (job or industry titles)

Logistic regression

Age, sex, interview method, cigarette smoking, body mass index, education, analgesics use

 

No response rates provided

 

Mellemgaar d et al., 1994

Cases, age 20–79 years, identified from the Danish Cancer Registry and pathology departments in 1989–1992 with histologic confirmation; controls selected from the Central Population Register and matched for age and sex

368 renal cell

396

Dry-cleaning work

Solvents

In-person interviews with questionnaire assessing most recent and longest-held occupation (job titles) and exposure to specific agents (self-reports)

Logistic regression

Age, body mass index, smoking

 

Response rates: 93.2% of cases, 85.6% of controls

 

Mandel et al., 1995

Cases, age 20–79 years, from six international sites, diagnosed and confirmed in 1989–1991 using cancer registries or surveillance of clinical and pathology departments; controls selected from population registers, electoral rolls, residential lists, HCFA records, or RDD, depending on the site; controls matched on age and sex

1732 renal

2309

Dry-cleaning solvents

Dry-cleaning work

In-person interviews to assess lifetime occupational history (job titles) and exposure to specific agents (self-reports)

Logistic regression

Age, center, body-mass index, cigarette smoking

 

No response rates provided

 

 

(Related to McCredie and Stewart, 1993)

 

Schlehofer et al., 1995

Cases, age 20–75 years, identified through 10 urology departments in the Rhein-Neckar-Odenwald area of Germany in 1989–1991 with histologic confirmation; controls randomly selected from population register and matched on age and sex

277 renal cell

286

Chlorinated solvents

In-person interview with questionnaire assessing exposure (in excess of 5 years) from list of specific substances (self-reports)

Logistic regression

 

 

Response rates: 97.3% of cases, 75% of controls

 

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Gérin et al., 1998

Male cases, age 35–75 years, diagnosed in one of 19 large Montreal-area hospitals in 1979–1985 and histologically confirmed; controls identified concurrently at 18 other cancer sites; age-matched, population-based controls were also chosen from electoral lists and random-digit dialing

177 kidney

1066, consisting of 533 population controls and 533 randomly selected subjects from other cases of cancer

Benzene

Toluene

Xylene

In-person interviews (direct or proxy) with segments on work histories (job titles and self-reported exposures); analyzed and coded by a team of chemists and industrial hygienists (about 300 exposures on semiquantitative scales)

Logistic regression

Age, family income, ethnicity, cigarette smoking, respondent status

 

Response rates: 82% of all cases, 71% of population controls

 

Vamvakas et al., 1998

Cases who underwent nephrectomy in 1987–1992 in a German hospital; controls selected from accident wards of three nearby hospitals

58 renal

84

Trichloroeth ylene

In-person interview (direct or proxy) with structured questionnaire assessing occupational history (job titles) and specific agent exposures (self-reports)

Logistic regression

Age, sex, smoking, body mass index, blood pressure, intake of diuretics

 

Response rates: 85% of cases, 75% of controls

 

Dosemeci et al., 1999

White cases, age 20–85 years, with histologically confirmed diagnosis identified through the Minnesota Cancer Surveillance System in 1988–1990; controls identified through RDD (age 21–64 years) and HCFA records (age 65–85 years), stratified for age and sex

438 renal cell carcinoma

687

Trichloroeth ylene

Tetrachloroe thylene

Solvents in general

In-person interviews (direct or proxy) with questionnaire assessing occupational history; job titles were coded and merged with a job-exposure matrix from NCI

Logistic regression

Age, smoking, hypertension, body mass index

 

Response rates: 87% of cases, 86% of controls

 

Pesch et al., 2000b

Cases in large hospitals in five regions in Germany in 1991–1995 with histologic confirmation; controls randomly selected from local residency registries matched on region, sex, and age.

935 renal-cell

4298

Trichloroeth ylene

Tetrachloroe thylene

Organic solvents

In-person interviews of lifetime occupational history using questionnaire to assess job titles and self-reported exposures; exposures ascertained by job-exposure matrixes

Condition al logistic regression

Matching variables, smoking

 

Response rates: 88% of cases, 71% of controls

 

NOTE: HCFA=Health Care Financing Administration; RDD=random-digit dialing.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

A small cohort study of German cardboard manufacturers exposed to trichloroethylene showed an increased risk of kidney cancer (SIR=7.97, 95% CI 2.59–8.59 and SMR=3.28, 95% CI=0.40–11.84) (Henschler et al., 1995). A case-control study using different cases from the same area as that investigated by Henschler and colleagues also showed an increased risk of kidney cancer (OR=10.8, 95% CI=3.36–34.75) (Vamvakas et al., 1998). A case-control study of kidney cancer in Minnesota showed an association only in women (OR=1.96, 95% CI=1.0–4.0) (Dosemeci et al., 1999). Pesch and colleagues (2000b) showed an increased risk of kidney cancer in men but not women with substantial exposure to trichloroethylene (OR=1.3, 95% CI =0.9–1.8; OR=0.8, 95% CI=0.3–1.9, respectively).

Some of the studies described above that investigated the risk of bladder cancer and exposure to tetrachloroethylene or dry-cleaning solvents also provided results on kidney cancer. The cohort studies of dry-cleaning workers had few exposed cases, and the associations had wide confidence intervals. Ruder and colleagues (2001) reported an increased risk of kidney cancer in dry-cleaning union workers (SMR=1.41, 95% CI=0.46–3.30) and in workers exposed to tetrachloroethylene only (SMR=1.73, 95% CI=0.21–6.25). The study by Blair and colleagues (1990) of dry-cleaning union members found no increased risk of kidney cancer (SMR=0.5, 95% CI=0.1–1.8), and Anttila and colleagues (1995) reported an increased risk of kidney cancer among workers occupationally exposed to tetrachloroethylene (SIR=1.82, 95% CI=0.22–6.56).

Pesch and colleagues (2000b) found an OR of 1.4 (95% CI=1.0–2.0) in men who had “substantial” exposure to tetrachloroethylene according to the German job-exposure matrix; the OR in women was 0.7 (95% CI=0.3–2.2). Mandel and colleagues (1995) also observed an increased risk of kidney cancer among those ever exposed to dry-cleaning solvents (OR=1.4, 95% CI=1.1–1.7). Both studies included large numbers of exposed cases, and the analyses were adjusted for smoking. Although a study of dry cleaners in Germany also reported an association with kidney cancer (OR=2.52, 95% CI=1.23–5.16), the authors indicated that subjects were exposed to a combination of tetrachloroethylene and tetrachlorocarbonate. The latter is a solvent not on the committee’s list to review, so this study (Schlehofer et al., 1995) was not considered critical to the committee’s review of tetrachloroethylene.

Other case-control studies reported associations with kidney cancer. Dosemeci and colleagues showed an OR of 1.12 (95% CI=0.7–1.7) in men exposed to tetrachloroethylene in Minnesota but an OR of 0.82 (95% CI=0.3–2.1) in women. A case-control study of dry cleaners in Denmark (Mellemgaard et al., 1994) showed an OR of 2.3 (95% CI=0.2–2.7) in men and an OR of 2.9 (95% CI=0.3–33) in women. A study in New South Wales (McCredie and Stewart, 1993) showed ORs of 2.49 (95% CI=0.97–6.35) for renal cancer and 4.68 (95% CI=1.32–16.56) for renal pelvic cancer. Another study of dry cleaners in Oklahoma (Asal et al., 1988) found an OR of 8.7 (95% CI=0.9–81.3) for kidney cancer. A study of residents in Cape Cod, Massachusetts (Aschengrau et al., 1993), reported an OR of 1.23 (95% CI=0.45–3.45) for kidney cancer after any exposure to tetrachloroethylene.

Gérin and colleagues (1998) investigated the risk of kidney cancer and exposure to benzene, toluene, and xylene; a weak association with no apparent exposure-response relationship was found for exposure to benzene (low exposure: OR=1.2, 95% CI=0.7–1.9; medium or high exposure: OR=1.3, 95% CI=0.7–2.4). No association was found for either toluene (medium or high exposure: OR=1.0, 95% CI=0.5–2.1) or xylene (medium or high exposure: OR=1.0, 95% CI=0.4–2.4).

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

The committee reviewed two other studies that investigated the association between kidney cancer and exposure to toluene. Wiebelt and Becker (1999) found no association (OR=0.49, 95% CI=0.06–2.34), and Walker and colleagues (1993) reported an increased risk in men who were potentially exposed to toluene in the shoe-manufacturing industry (SMR=1.71, 95% CI=0.62–3.73).

A cohort study of US isopropyl manufacturers found an association (SMR=6.45, 95% CI=0.78–23.29) between exposure to isopropyl alcohol and kidney cancer (Alderson and Rattan, 1980).

The committee reviewed several cohort studies of painters and workers in aircraft manufacturing (Boice et al., 1999; Garabrant et al., 1988), shoe manufacturing (Fu et al., 1996; Walker et al., 1993), transformer assembly (Greenland et al., 1994), and petroleum refining (Poole et al., 1993) and several case-control studies conducted in Minnesota (Dosemeci et al., 1999), Denmark (Mellemgaard et al., 1994), New South Wales (McCredie and Stewart, 1993), Finland (Partanen et al., 1991), and the United Kingdom (Harrington et al., 1989).

The largest cohort study of painters showed no association between being a painter and kidney cancer (SMR=1.06, 95% CI=0.86–1.29) (Steenland and Palu, 1999), but a positive association was found in other cohorts of painters, including a study by Matanoski and colleagues (1986) (SMR=1.28, 95% CI=0.91–1.76). A study of paint-manufacturing workers showed no association with kidney cancer (SMR=0.39) (Morgan et al., 1981). Several case-control studies that considered occupation as a painter or in paint manufacturing reported positive associations with kidney cancer (Asal et al., 1988: OR=1.3, 95% CI=0.7–2.6; Jensen et al., 1988: OR=1.8, 95% CI=0.7–4.6). Pesch and colleagues (2000b) found that relative risks increased with increasing duration of exposure (medium: OR=1.6, 95% CI=0.8–3.0; very long: OR=2.3, 95% CI 0.8–6.8).

No evidence of an association between exposure to unspecified mixtures of solvents and kidney cancer was found in two cohorts of aircraft manufacturers in California (Boice et al., 1999: SMR=0.81, 95% CI=0.44–1.36; Garabrant et al., 1988: SMR=0.93, 95% CI=0.48–1.64), and the association decreased with increasing years of exposure (Boice et al., 1999). No association was found in an incidence study of Finnish workers monitored for halogenated hydrocarbon exposure (SIR=0.89, 95% CI=0.36–1.82) (Anttila et al., 1995) or among US petroleum-refinery workers exposed to aromatic hydrocarbons (OR=0.95, 95% CI=0.50–1.80). However, a study of dry cleaners in Germany exposed to a mixture of solvents, including tetrachloroethylene and tetrachlorocarbonate, reported an association with kidney cancer (OR=2.52, 95% CI=1.23–5.16) (Schlehofer et al., 1995), as did a study of US transformer-assembly workers exposed to solvents (OR=1.64, 95% CI=0.49–5.50) (Greenland et al., 1994). Two cohort studies of shoe-manufacturing workers also showed an increased association, including Fu and colleagues’ (1996) study of Florence workers (SMR=4.00, 95% CI=0.83–11.69 for high solvent exposure) and Walker and colleagues’ (1993) study (SMR=1.71, 95% CI=0.62–3.73 among men; the SMR was 0.97 among women).

Case-control studies that analyzed self-reported exposure to organic solvents include studies conducted in Minnesota (Dosemeci et al., 1999) and Denmark (Mellemgaard et al., 1994). The study in Minnesota showed ORs close to 1.0: in men, it was 1.12 (95% CI=0.7–1.7), and in women it was 0.82 (95% CI=0.3–2.1). The study in Denmark showed increased risks in both men (OR=2.3, 95% CI=0.2–27) and women (OR=2.9, 95% CI=0.3–33). Other case-control studies reported an association between kidney cancer and nonchlorinated solvents (Partanen et al., 1991: OR=3.46, 95% CI=0.91–13.2) and organic solvents (Sharpe et al.,

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

1989: OR=1.68, 95% CI=0.89–3.18). Pesch and colleagues (2000b) found positive associations between exposure to organic solvents specifically, using a job-exposure matrix, and to solvents broadly, using a job task-exposure matrix. The relative risks in men were increased on the basis of the job-exposure matrix (OR=1.6, 95% CI=1.1–2.3) and the job task-exposure matrix (OR=1.5, 95% CI=1.0–2.3). The risks were increased in women on the basis of the job task-exposure matrix (OR=2.1, 95% CI=1.0–4.4), but not the job-exposure matrix (OR=0.3, 95% CI=0.1–1.3). No association was found in the study by Harrington and colleagues (1989) that examined exposure to solvents in the UK (OR=1.0, 95% CI=0.2–4.9).

Summary and Conclusions

Because of their small numbers of cases, most studies on exposure to trichloroethylene and kidney cancer lacked the power to detect excess risks. Although positive associations were suggested by three studies (Henschler et al., 1995; Pesch et al., 2000b; Vamvakas et al., 1998), one was based on self-reported exposures (Vamvakas et al., 1998), and the committee was concerned about bias. The committee did not find the results of the other case-control study to be persuasive (Pesch et al., 2000b). The findings of the cohort study (Henschler et al., 1995) were based on a cluster of cases. For exposure to benzene, toluene, xylene, and isopropyl alcohol, there was only a small number of studies available, and they lacked consistently positive findings. Therefore, the committee determined that the evidence for these exposes was insufficient to determine whether an association exists for kidney cancer.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to solvents under review, other than tetrachloroethylene and dry-cleaning solvents, and kidney cancer.

For exposure to tetrachloroethylene and dry-cleaning solvents, positive associations were observed in several well-conducted studies (Mandel et al., 1995; McCredie and Stewart, 1993; Pesch et al., 2000b), and the committee determined that the evidence between exposure to tetrachloroethylene and dry-cleaning solvents and risk of kidney cancer was limited/suggestive of an association.

The committee concludes, from its assessment of the epidemiologic literature, that there is limited/suggestive evidence of an association between chronic exposure to tetrachloroethylene and dry-cleaning solvents and kidney cancer.

Although several case-control studies showed an increased risk of kidney cancer with exposure to unspecified mixtures of organic solvents (including Dosemeci et al., 1999; Jensen et al., 1988; Mellemgaard et al., 1994; Partanen et al., 1991; Pesch et al., 2000b; and Sharpe et al., 1989), several cohort studies did not, including the largest study of painters (Steenland and Palu, 1999), the painter study by Morgan and colleagues (1981), two studies of aircraft manufacturers (Boice et al., 1999; Garabrant et al., 1988), a biologically monitored study (Anttila et al., 1995), and a study on petroleum-refinery workers (Poole et al., 1993). As a result, the committee was unable to reach a consensus on an association between kidney cancer and exposure to organic solvents. Some committee members believed that the evidence was limited/suggestive, and others believed that it was inadequate/insufficient to determine whether an association exists. In evaluating each study and the overall body of evidence, committee members differed in their judgment about the extent to which bias and confounding affected the results.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

The large number of positive findings from the case-control studies, especially among women, and the use of a job-exposure matrix by Pesch and colleagues (2000b) were supportive of a conclusion that the evidence was limited/suggestive of an association between exposure to unspecified mixtures of solvents and kidney cancer. In contrast, the lack of positive findings in large cohorts of occupationally exposed populations—such as painters, aircraft manufacturers, and petroleum refinery workers—and in a biologically monitored cohort of Finnish workers (Anttila et al., 1995), the lack of increased risks as exposure and duration of exposure increased (Boice et al., 1999), and the large number of case-control studies that relied on self-reported exposures (Asal et al., 1988; Dosemeci et al., 1999; Jensen et al., 1988; McCredie and Stewart, 1993; Mellemgaard et al., 1994; Schlehofer et al., 1995; Sharpe et al., 1989) supported a conclusion that the evidence was inadequate/insufficient to determine whether an association exists.

Thus, the committee could not reach a consensus conclusion for kidney cancer and exposure to unspecified mixtures of organic solvents. As more studies are conducted on organic solvents and the risk of kidney cancer, future committees may revisit this literature in evaluating the evidence of association.

The studies reviewed by the committee in making its conclusions are identified below in Table 6.30. Unless indicated in the table, the study populations include both men and women.

TABLE 6.30 Selected Epidemiologic Studies—Kidney Cancer and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Studies—Incidence

Hansen et al., 2001

Biologically monitored Danish workers

 

Males, ever exposed

3

0.9 (0.2–2.6)

 

Females, ever exposed

1

2.4 (0.03–14)

Blair et al., 1998

Aircraft-maintenance workers in Utah

 

 

No exposure

9

1.6 (0.5–5.4)

 

<5 unit-years

9

1.4 (0.4–4.7)

 

5–25 unit-years

5

1.3 (0.3–4.7)

Henschler et al., 1995

Male German cardboard manufacturers, employed >1 year

5

7.97 (2.59–8.59)

Anttila et al., 1995

Finnish workers occupationally exposed

 

Entire period since first measurement

6

0.87 (0.32–1.89)

 

0–9

1

0.53 (0.01–2.95)

 

10–19

5

1.39 (0.45–3.24)

 

20+

0

—(0.00–2.48)

Axelson et al., 1994

Swedish men occupationally exposed

6

1.16 (0.42–2.52)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California, routine exposure

7

0.99 (0.40–2.04)

Blair et al., 1998

Aircraft-maintenance workers in Utah

 

 

No trichloroethylene exposure

10

2.5 (0.7–8.9)

 

<5 unit-years

8

2.0 (0.5–7.6)

 

5–25 unit-years

1

0.4 (0.1–4.0)

 

Low-level intermittent

12

2.1 (0.6–7.5)

 

Low-level continuous

9

2.2 (0.6–8.1)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Morgan et al., 1998

Aerospace workers in Arizona

 

Any exposure

8

1.32 (0.57–2.60)

 

Low exposure

1

0.47 (0.01–2.62)

 

High exposure

7

1.78 (0.72–3.66)

Henschler et al., 1995

Male German cardboard manufacturers, employed >1 year

2

3.28 (0.40–11.84)

Greenland et al., 1994

White male US transformer manufacturers, ever exposed

NA

0.99 (0.30–3.32)

Case-Control Studies

Pesch et al., 2000b

Participants in multiple centers in Germany

 

German job-exposure matrix

 

 

Trichloroethylene (males)

 

 

Medium

135

1.1 (0.9–1.4)

 

High

138

1.1 (0.9–1.4)

 

Substantial

55

1.3 (0.9–1.8)

 

Trichloroethylene (females)

 

 

Medium

28

1.2 (0.8–1.8)

 

High

29

1.3 (0.8–2.0)

 

Substantial

6

0.8 (0.3–1.9)

 

Job task-exposure matrix approach

 

 

Trichloroethylene (males)

 

 

Medium

68

1.3 (1.0–1.8)

 

High

59

1.1 (0.8–1.5)

 

Substantial

22

1.3 (0.8–2.1)

 

Trichloroethylene (females)

 

 

Medium

11

1.3 (0.7–2.6)

 

High

7

0.8 (0.4–1.9)

 

Substantial

5

1.8 (0.6–5.0)

Dosemeci et al., 1999

Residents of Minnesota

55

1.30 (0.9–1.9)

Males

33

1.04 (0.6–1.7)

 

Females

22

1.96 (1.0–4.0)

Vamvakas et al., 1998

Residents of Germany with long-term exposure

19

10.8 (3.36–34.75)

Tetrachloroethylene and Dry-cleaning Solvents

Cohort Study—Incidence

Anttila et al., 1995

Finnish workers monitored for exposure

2

1.82 (0.22–6.56)

Cohort Studies—Mortality

Ruder et al., 2001

Dry-cleaning labor-union workers

5

1.41 (0.46–3.30)

Tetrachloroethylene only

2

1.73 (0.21–6.25)

 

Tetrachloroethylene plus other solvents (likely Stoddard)

3

1.27 (0.26–3.72)

Blair et al., 1990

Members of a dry-cleaning union in St. Louis, MO

2

0.5 (0.1–1.8)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Case-Control Studies

Pesch et al., 2000b

Participants in multiple centers in Germany

 

German job-exposure matrix

 

 

Tetrachloroethylene (males)

 

 

Medium

154

1.4 (1.1–1.7)

 

High

119

1.1 (0.9–1.4)

 

Substantial

50

1.4 (1.0–2.0)

 

Tetrachloroethylene (females)

 

 

Medium

12

0.7 (0.4–1.3)

 

High

19

1.1 (0.7–1.9)

 

Substantial

4

0.7 (0.3–2.2)

 

Job task-exposure matrix approach

 

 

Tetrachloroethylene (males)

 

 

Medium

44

1.2 (0.9–1.7)

 

High

39

1.1 (0.7–1.5)

 

Substantial

15

1.3 (0.7–2.3)

 

Tetrachloroethylene (females)

 

 

Medium

8

2.2 (0.9–5.2)

 

High

6

1.5 (0.6–3.8)

 

Substantial

3

2.0 (0.5–7.8)

Dosemeci et al., 1999

Residents of Minnesota

50

1.07 (0.7–1.6)

Males, ever exposed

42

1.12 (0.7–1.7)

 

Females, ever exposed

8

0.82 (0.3–2.1)

Schlehofer et al., 1995

Residents of Germany

 

Tetrachloroethylene and tetrachlorocarbonate, exposed >5 years

27

2.52 (1.23–5.16)

Mandel et al., 1995

International renal cell carcinoma cases

 

Dry-cleaning industry, ever employed

8

0.9 (0.3–2.4)

 

Dry-cleaning solvents, ever exposed

245

1.4 (1.1–1.7)

 

1–7 years

70

1.2 (0.9–1.8)

 

8–25 years

98

1.7 (1.2–2.4)

 

26–60 years

75

1.2 (0.9–1.8)

Mellemgaard et al., 1994

Residents of Denmark, employed >10 years before as dry cleaners

 

 

Males

2

2.3 (0.2–27)

 

Females

2

2.9 (0.3–33)

McCredie and Stewart, 1993

Residents of New South Wales, ever employed in dry-cleaning industry

 

 

Renal

16

2.49 (0.97–6.35)

 

Renal pelvis

8

4.68 (1.32–16.56)

Aschengrau et al., 1993

Residents of upper Cape Cod, MA

 

Any exposure

6

1.23 (0.40–3.11)

 

Low exposure

6

1.36 (0.45–3.45)

Asal et al., 1988

Female residents of Oklahoma, usual dry-cleaning occupation

NA

8.7 (0.9–81.3)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Benzene

Case-Control Study

Gérm et al., 1998

Male residents of Montreal, Canada

 

Low exposure

27

1.2 (0.7–1.9)

 

Medium or high exposure

12

1.3 (0.7–2.4)

Toluene

Cohort Studies—Mortality

Wiebelt and Becker, 1999

Male German rotogravure printers, employed >1 year

2

0.49 (0.06–2.34)

Walker et al., 1993

Shoe manufacturers in two plants in Ohio, potentially exposed to toluene and other solvents

 

 

Total

9

1.36 (0.62–2.59)

 

Males

6

1.71 (0.62–3.73)

 

Females

3

0.97 (0.20–2.84)

Case-Control Study

Gérin et al., 1998

Male residents of Montreal, Canada

 

Low exposure

17

0.9 (0.5–1.5)

 

Medium or high exposure

8

1.0 (0.5–2.1)

Xylene

Case-Control Study

Gérin et al., 1998

Male residents of Montreal, Canada

 

Low exposure

17

1.0 (0.6–1.7)

 

Medium or high exposure

6

1.0 (0.4–2.4)

Isopropyl Alcohol

Cohort Study—Mortality

Alderson and Rattan, 1980

Male UK isopropyl manufacturers, employed >1 year

2

6.45 (0.78–23.29)a

Unspecified Mixtures of Organic Solvents

Cohort Study—Incidence

Anttila et al., 1995

Finnish workers monitored for exposure

 

Halogenated hydrocarbons

7

0.89 (0.36–1.82)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

Mixed solvents, routine exposure

14

0.81 (0.44–1.36)

 

Years exposed

 

 

<1

11

1.54 (0.74–3.20)

 

1–4

19

1.08 (0.58–2.01)

 

≥5

14

0.41 (0.20–0.82)

Steenland and Palu, 1999

US painters and other union members

 

Painters

100

1.06 (0.86–1.29)

 

20 years since membership

71

0.97 (0.76–1.22)

 

Nonpainters

21

0.77 (0.48–1.18)

 

20 years since membership

12

0.67 (0.34–1.17)

Fu et al., 1996

Shoe-manufacturing workers

 

 

English workers:

 

 

Probable solvent exposure

1

0.25 (0.01–1.37)

 

High solvent exposure

0

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

 

Florence workers:

 

 

Probable solvent exposure

3

3.53 (0.73–10.31)

 

High solvent exposure

3

4.00 (0.83–11.69)

Greenland et al., 1994

White male US transformer-assembly workers

 

Solvents, ever exposed

NA

1.64 (0.49–5.50)

Poole et al., 1993

Male US petroleum-refinery workers

 

Aromatic hydrocarbons, ever exposed

80

0.95 (0.50–1.80)

 

Chlorinated solvents, ever exposed

12

0.69 (0.32–1.50)

Walker et al., 1993

Shoe-manufacturing workers in Ohio, potentially exposed

9

1.36 (0.62–2.59)

 

Males

6

1.71 (0.62–3.73)

 

Females

3

0.97 (0.20–2.84)

Garabrant et al., 1988

Aircraft-manufacturing workers in California

12

0.93 (0.48–1.64)

Matanoski et al., 1986

US painters and allied tradesmen, union members

38

1.28 (0.91–1.76)

Morgan et al., 1981

Male US paint and coatings manufacturers, employed >1 year

5

0.39

Case-Control Studies

Pesch et al., 2000b

Residents of multiple centers in Germany

 

Painters or dyers, males, duration of exposure:

 

 

Medium

12

1.6 (0.8–3.0)

 

Long

10

1.4 (0.7–2.8)

 

Very long

5

2.3 (0.8–6.8)

 

Job-exposure matrix, organic solvents:

 

 

Males, substantial exposure

38

1.6 (1.1–2.3)

 

Females, substantial exposure

2

0.3 (0.1–1.3)

 

Job task-exposure matrix, solvents:

 

 

Males, substantial exposure

33

1.5 (1.0–2.3)

 

Females, substantial exposure

10

2.1 (1.0–4.4)

Dosemeci et al., 1999

Residents of Minnesota

 

Solvents, ever exposed

126

1.16 (0.9–1.5)

 

Males

91

0.93 (0.7–1.3)

 

Females

35

2.29 (1.3–4.2)

Schlehofer et al., 1995

Residents of Germany

 

Tetrachloroethylene and tetrachlorocarbonate, exposed >5 years

27

2.52 (1.23–5.16)

Mellemgaard et al., 1994

Residents of Denmark

 

Solvents, exposed >10 years before

 

 

Males

50

1.5 (0.9–2.4)

 

Females

16

6.4 (1.8–23.0)

McCredie and Stewart, 1993

Residents of New South Wales, ever exposed to solvents

 

 

Renal

109

1.54 (1.11–2.14)

 

Renal pelvis

24

1.40 (0.82–2.40)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Partanen et al., 1991

Residents of Finland

 

Nonchlorinated solvents, males exposed >5 years

9

3.46 (0.91–13.2)

Harrington et al., 1989

Residents of the UK

 

Solvents, ever exposed

8

1.0 (0.2–4.9)

Sharpe et al., 1989

Residents of Canada

 

Organic solvents, regular exposure

33

1.68 (0.89–3.18)

Asal et al., 1988

Residents of Oklahoma

 

 

Painting or paint manufacturing, usual occupation

22

1.3 (0.7–2.6)

Jensen et al., 1988

Residents of Denmark

 

Painter or paint manufacturing, ever employed

10

1.8 (0.7–4.6)

NOTE: NA=not available.

a95% CI calculated by the committee with standard methods from the observed and expected numbers presented in the original study.

BRAIN AND OTHER CENTRAL NERVOUS SYSTEM CANCERS

Description of Case-Control Studies

Three case-control studies of brain cancer were identified. They included different histologic types of brain cancer: astrocytic (Heineman et al., 1994; Thomas et al., 1987), gliomas (Rodvall et al., 1996), and unspecified brain cancers (Paulu et al., 1999). Occupational exposure information was obtained from next of kin (Heineman et al., 1994; Thomas et al., 1987) or from subjects (Rodvall et al., 1996), and in a study of tetrachloroethylene in drinking water the exposure information was based on water quality data (Paulu et al., 1999). In two studies, industrial hygienists attributed exposure to specific solvents on the basis of their own expertise and information provided by subjects (Rodvall et al., 1996; Thomas et al., 1987), and a job-exposure matrix was used in an updated analysis of the Thomas and colleagues’ study (1987) (Heineman et al., 1994) (see Table 6.31).

Epidemiologic Studies of Exposure to Organic Solvents and Brain and Central Nervous System Cancers

Most studies found null associations when investigating the relationship between trichloroethylene and risk of brain and other CNS cancers, including studies of workers biologically monitored for exposure (SIR=1.09, 95% CI=0.50–2.07) (Anttila et al., 1995), of workers in the aircraft and aerospace industries (Blair et al., 1998: SMR=0.8, 95% CI=0.2–2.2; Boice et al., 1999: SMR=0.54, 95% CI=0.15–1.37; Morgan et al., 1998: SMR=0.55, 95% CI=0.15–1.40), and of workers in transformer assembly (OR=0.93, 95% CI=0.32–2.69) (Greenland et al., 1994).

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.31 Description of Case-Control Studies of Brain and Central Nervous System Cancers and Exposure to Organic Solvents

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Thomas et al., 1987

White, male cases, age 30 years or more, identified from death certificates in southern Louisiana, northern New Jersey, and Philadelphia in 1978–1981; deceased controls from other causes collected in the same fashion, matched for age, year of death, and study area (see also Heineman et al., 1994)

300 astrocytic brain tumors

386

Organic solvents

In-person interview of next of kin with structured questionnaire to assess lifetime occupational history (job titles); probability of exposure determined by industrial hygienist assessment

Maximum likelihood estimate

Age at death, cigarette smoking, alcoholism, ethnicity, education

 

Response rates (next-of-kin): 74% of cases, 63% of controls

 

Heineman et al., 1994

See above, Thomas et al., 1987

Same cases as above

320

Trichloroethylene

Tetrachloroethylene

Methylene chloride

Chloroform

Organic solvents

In-person interview of next of kin with structured questionnaire to assess lifetime occupational history (job titles); probability of exposure to six chlorinated hydrocarbons determined with job-exposure matrixes

Logistic regression

Age, study area

 

Response rates (next-of-kin): 74% of cases, 63% of controls

 

Rodvall et al., 1996

Cases, age 25–74 years, reported to the Regional Cancer Registry among residents of the catchment area of the Uppsala University Hospital in Sweden, with histologic confirmation; controls selected from parish records and matched for sex and age

151 brain gliomas

343

Trichloroethylene

Benzene

Toluene

Xylene

Solvents

Self-reported occupational history (job titles) and exposure to specific agents (self-reports) reviewed by occupational hygienist and to assess probability of exposure to specific agents

Multiple logistic regression

Sex, age, population density

 

Response rates: 79% of cases, 82% of controls

 

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Paulu et al., 1999

Cases reported to the Massachusetts Cancer Registry, diagnosed in 1983–1986 among residents of five upper Cape Cod towns; living controls were selected from the records of HCFA and through RDD; deceased controls identified by the state Department of Vital Statistics and Research files

36 brain

703

Tetrachloroethylene

Exposure dose estimated in areas of contaminated drinking water, accounting for location and years of residence, water flow, pipe characteristics

Multiple logistic regression

Age at diagnosis, vital status, sex, occupational exposure to solvents; specific cancer risk factors controlled in respective analyses

 

Response rates: 79% of cases, 76% of HCFA controls, 74% of RDD controls, 79% of next of kin of deceased controls

 

NOTE: HCFA=Health Care Financing Administration; RDD=random-digit dialing.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Three studies reported positive associations. Ritz (1999) found a 27% increase in mortality in US uranium processing plant workers (SMR=1.27, 95% CI=0.66–2.22) and found that relative risks increased with duration as well as with level of exposure to trichloroethylene. Heineman and colleagues (1994), in a case-control study of astrocytic brain cancer, found no increases in relative risk of brain cancer (OR=1.1, 95% CI=0.8–1.6), although relative risks increased with duration of exposure among subjects whose exposures were classified as highly probable (2–20 years: OR=1.1, 95% CI=0.3–3.7; more than 20 years: OR=6.1, 95% CI=0.7–143.5). Rodvall and colleagues (1996) found an increased risk of glioma among men who reported exposure to trichloroethylene (OR=2.4, 95% CI=0.9–6.4).

No associations between exposure to tetrachloroethylene or to dry-cleaning solvents and brain or other CNS cancers were found in the three cohort studies reviewed by the committee (Anttila et al., 1995; Blair et al., 1990; Boice et al., 1999). No association was found in the case-control study of contaminated drinking water (Paulu et al., 1999). A 20% increased risk (OR=1.2, 95% CI=0.8–1.6) was found in the case-control study by Heineman and colleagues (1994), but the relative risks did not increase with increasing probability of exposure.

Two studies of photographic film-base manufacturing workers showed an increased risk of brain and other CNS cancer deaths (Hearne and Pifer, 1999: SMR=2.16, 95% CI=0.79–4.69; Tomensen et al., 1997: SMR=1.45, 95% CI=0.40–3.72). In a study of two nuclear facilities, Carpenter and colleagues (1988) found that the risk of CNS cancer increased with increasing years of exposure to methylene chloride (10–20 years of exposure, 10-year latency: OR=1.8; over 20 years of exposure, 10-year latency: OR=4.01), but the numbers of cases were too small to form any meaningful inferences. In contrast, Heineman and colleagues (1994) found that the risk of astrocytic brain cancer increased with the probability of exposure (medium probability: OR=1.6, 95% CI=0.8–3.0; high probability: OR=2.4, 95% CI=1.0–5.9) and that risk increased with duration of exposure among subjects with high probability of exposure to methylene chloride (2–20 years: OR=1.8, 95% CI=0.6–6.0; over 20 years: OR=6.1, 95% CI 1.1–43.8).

Yin and colleagues (1996a) studied benzene-exposed workers in China and reported a 30% increase in the risk of malignant brain tumors and benign tumors of the brain and unspecified CNS (RR=1.3, 95% CI=0.5–4.1). When levels of exposure to benzene were incorporated into a second analysis (Hayes et al., 1996), risk did not increase with increasing cumulative exposure. In a cohort study of women in Shanghai, China, with high probability of exposure to benzene, Heineman and colleagues (1995) found an 80% excess incidence of brain cancer (SIR=1.8, 95% CI=1.1–2.8); they also found a dose-response relationship between low exposure to benzene and high exposure (SIR=1.5, 95% CI=0.6–3.4 for low, and SIR=1.8, 95% CI=1.1–2.9 for high). In a cohort of benzene-exposed workers at Monsanto chemical plants, risk among production workers increased with increasing cumulative exposure, and maintenance workers showed a 20% excess risk (SMR=1.2, 95% CI=0.4–2.5) (Ireland et al., 1997).

A nested case-control study of transformer-assembly workers showed an increased risk of malignant and unspecified brain tumors (OR=2.1, 95% CI=1.0–4.4) (Greenland et al., 1994). The nested case-control study of workers at two nuclear facilities showed no increase in CNS tumors among those exposed to benzene (ever exposed, 10-year latency: OR=0.57) (Carpenter et al., 1988). In a population-based case-control study of brain gliomas in Sweden (Rodvall et al., 1996), a 450% increase in relative risk was found (OR=5.5, 95% CI=1.4–21.3).

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Only a few investigations have examined the relationship between brain and other CNS cancers and other solvents, including xylene, toluene (Anttila et al., 1998; Rodvall et al., 1996), phenol (Dosemeci et al., 1991), diethylene glycol, ethanol, isopropanol (Leffingwell et al., 1983), and chloroform (Heineman et al., 1994). They yielded no strong support of associations with any of these solvents.

Many of the previously described studies also examined the risk of brain and other CNS cancers in relation to unspecified mixtures of organic solvents. Of the cohort studies, only Boice and colleagues (1999) did not show an increased relative risk among aircraft-manufacturing workers in California who were routinely exposed to mixed solvents (SMR=0.68, 95% CI=0.36–1.16). The study of women in Shanghai, China, showed an SIR of 2.0 (95% CI=1.3–3.0) with high probability of exposure to solvent mixtures (Heineman et al., 1995), which was supported by a study of US transformer-assembly workers who were ever exposed to solvents (OR=2.65, 95% CI=0.84–8.36) (Greenland et al., 1994). On the basis of probability of exposure to solvents, Heineman and colleagues (1995) observed a dose-response relationship with an SIR of 1.2 (95% CI=1.0–1.5) for low probability of exposure and 2.0 (95% CI=1.3–3.0) for high probability. Carpenter and colleagues (1988) also showed an increased relative risk among workers at two nuclear facilities exposed to various solvents (exposed to trichloroethylene, tetrachloroethylene, and methyl chloroform, no latency: OR=1.76; with 10-year latency: OR=1.26). Those exposed to toluene, xylene, and methyl ethyl ketone showed an OR of 1.96 with no latency and 1.37 with 10-year latency.

Two case-control studies also showed positive associations between exposure to unspecified mixtures of organic solvents and brain and other CNS cancer. The study by Rodvall and colleagues (1996) showed an OR of 2.6 (95% CI=1.3–5.2) for brain gliomas among men with self-reported exposure to solvents; the OR for women was 0.4 (95% CI=0.1–2.0). Among people with astrocytic brain tumors in Louisiana, New Jersey, and Philadelphia, the odds ratios increased with increasing probability of exposure (low probability: OR=1.1, 95% CI=0.6–1.7; medium probability: OR=1.5, 95% CI=0.8–2.7; high probability: OR=1.4, 95% CI=0.9–2.1) and with duration of employment (2–20 years: OR=1.1, 95% CI=0.7–1.7; over 20 years: OR=1.7, 95% CI=1.1–2.6).

Summary and Conclusion

Although several studies on trichloroethylene exposure and brain and CNS cancers were positive, a number of limitations weakened their support for an association. Of particular concern was the study by Ritz (1999) of uranium processing plant workers who were also exposed to radioactive dust that could have been associated with brain cancer. In addition, most studies had wide CIs associated with the relative risk estimates, and there was a lack of a body of evidence related to specific types of brain cancer. For exposure to tetrachloroethylene and dry-cleaning solvents, most of the studies did not find an association. In contrast, one study on methylene chloride showed a strong association with astrocytic brain cancer (Heineman et al., 1994) but it was not supported by findings from the cohort studies of all types of brain and other CNS cancers.

Two cohort studies (Greenland et al., 1994; Heineman et al., 1995) and one population-based study in Sweden (Rodvall et al., 1996) of brain cancer showed increased risks with exposure to benzene. In the cohort studies, all brain cancers were grouped together; in the case-control study, histologically confirmed cases of glioma were included (Rodvall et al., 1996). There is no likelihood of recall bias in the case-control study by Rodvall and colleagues,

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

inasmuch as exposures were attributed by industrial hygienists unaware of the status of subjects. The committee did not believe that any of these studies were subject to strong confounding biases, in that risk factors for brain and other CNS cancer are not well known. A dose-response relationship between brain and other CNS cancers and increasing levels of exposure was supported by two studies: Heineman et al., 1995; Ireland et al., 1997. However, the study by Hayes and colleagues (1996) did not yield a dose-response relationship based on years of exposure to benzene, nor did the nested case-control study by Carpenter and colleagues (1988) support an association with brain or other CNS cancers. Moreover, several studies were based on small numbers of exposed cases. As a result, the committee did not reach a consensus. Some committee members believed that the evidence was limited/suggestive of an association; others believed it was inadequate/insufficient to determine whether an association exists. Additional research will help to clarify whether an association exists between exposure to benzene and the risk of brain and other CNS cancers.

All but one of the studies reviewed by the committee reported positive associations between exposure to solvent mixtures and brain and other CNS cancers, including two that showed a slight dose-response pattern (Heineman et al., 1994, 1995). However, the studies did not examine the same cancer outcomes; some evaluated only brain cancer, others looked at brain and other CNS cancers together, and still others evaluated specific subtypes of brain cancer. Furthermore, the mixtures of solvents were known in some studies and unspecified in others. Given those concerns and the consistently positive findings, the committee could not reach a consensus; the evidence was neither wholly inadequate/insufficient to determine whether an association exists nor wholly limited/suggestive.

Table 6.32 identifies the key studies and relevant data points reviewed by the committee in drawing its conclusions. Unless indicated in the tables, the study populations include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to the solvents under review, other than benzene, and brain or other central nervous system cancers.

TABLE 6.32 Selected Epidemiologic Studies—Brain and Central Nervous System Tumors and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Study—Incidence

Anttila et al., 1995

Biologically monitored workers in Finland

 

 

Entire period since first measurement

9

1.09 (0.50–2.07)

 

0–9 years

0

—(0–1.26)

 

10–19 years

8

2.00 (0.86–3.93)

 

20+ years

1

0.76 (0.02–4.26)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California, potential routine exposure

4

0.54 (0.15–1.37)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Ritz, 1999

White male US uranium-processing plant workers

 

 

Trichloroethylene, cutting fluids, or kerosene

12

1.27 (0.66–2.22)

 

Trichloroethylene—light exposure

 

 

>2 years, no latency

6

1.81 (0.49–6.71)

 

>2 years, 15-year latency

4

2.29 (0.42–12.5)

 

>5 years, no latency

3

1.32 (0.28–6.17)

 

>5 years, 15-year latency

3

5.41 (0.87–33.9)

 

Trichloroethylene—moderate exposure

 

 

>2 years, no latency

1

3.26 (0.37–28.9)

 

>2 years, 15-year latency

1

6.94 (0.66–73.1)

 

>5 years, no latency

1

4.52 (0.49–41.5)

 

>5 years, 15-year latency

1

14.4 (1.24–167.0)

Blair et al., 1998

Male aircraft-maintenance workers in Utah

11a

0.8 (0.2–2.2)

 

<5 unit-years

3

2.0 (0.2–19.7)

 

5–25 unit-years

4

3.9 (0.4–34.9)

 

>25 unit-years

1

0.8 (0.1–13.2)

Morgan et al., 1998

Aerospace workers in Arizona, employed >6 months

4

0.55 (0.15–1.40)

Greenland et al., 1994

White male US transformer-assembly workers, ever exposed

NA

0.93 (0.32–2.69)

Case-Control Studies

Rodvall et al., 1996

Hospital catchment-area residents of Sweden, ever exposed (gliomas)

NA

2.4 (0.9–6.4)

Heineman et al., 1994

Male residents of Louisiana, New Jersey, and Philadelphia (astrocytic brain tumors)

 

 

Ever exposed

128

1.1 (0.8–1.6)

 

Low probability

67

1.1 (0.7–1.7)

 

Medium probability

42

1.1 (0.6–1.8)

 

High probability

12

1.1 (0.5–2.8)

 

Duration of employment

 

 

2–20 years

7

1.1 (0.3–3.7)

 

21+ years

5

6.1 (0.7–143.5)

Tetrachloroethylene and Dry-cleaning Solvents

Cohort Study—Incidence

Anttila et al., 1995

Biologically monitored workers in Finland, ever exposed

2

1.15 (0.14–4.15)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California, potential routine exposure

2

0.59 (0.07–2.14)

Blair et al., 1990

Dry-cleaning union members in Missouri

1

0.2 (0.0–1.2)

Case-Control Studies

Paulu et al., 1999

Massachusetts residents exposed through public drinking water

3

0.6 (0.1–1.7)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Heineman et al., 1994

Male residents of Louisiana, New Jersey, and Philadelphia (astrocytic brain tumors)

 

 

Ever exposed

111

1.2 (0.8–1.6)

 

Low probability

72

1.3 (0.8–1.9)

 

Medium probability

30

0.9 (0.5–1.6)

 

High probability

9

1.2 (0.4–3.5)

 

Duration of employment

 

 

2–20 years

71

1.1 (0.7–1.6)

 

21+ years

28

1.4 (0.7–2.7)

Methylene Chloride

Cohort Studies—Mortality

Hearne and Pifer, 1999

Male US photographic-film base manufacturing workers, employed >1 year

6

2.16 (0.79–4.69)

Tomenson et al., 1997

Male UK photographic-film base manufacturing workers, ever employed

4

1.45 (0.40–3.72)

Carpenter et al., 1988

Workers at two nuclear facilities

 

 

All workers

 

 

No latency

29

1.38

 

10-year latency

21

1.00

 

Workers with moderate or high exposure:

 

 

>1 and ≤3 years of exposure

 

 

No latency

2

1.19

 

10-year latency

0

 

>3 and ≤10 years of exposure

 

 

No latency

3

1.33

 

10-year latency

2

0.83

 

>10 and ≤20 years of exposure

 

 

No latency

1

0.50

 

10-year latency

1

1.80

 

>20 years of exposure

 

 

No latency

1

4.00

 

10-year latency

1

4.01

Case-Control Study

Heineman et al., 1994

Male residents of Louisiana, New Jersey, and Philadelphia (astrocytic brain tumors)

 

 

Ever exposed

119

1.3 (0.9–1.8)

 

Low probability

71

1.0 (0.7–1.6)

 

Medium probability

29

1.6 (0.8–3.0)

 

High probability

19

2.4 (1.0–5.9)

 

Duration of employment

 

 

2–20 years

9

1.8 (0.6–6.0)

 

21+ years

8

6.1 (1.1–43.8)

Benzene

Cohort Study—Incidence

Heineman et al., 1995

Women in Shanghai, China

 

 

Low exposure probability

5

1.6 (0.5–3.7)

 

High exposure probability

19

1.8 (1.1–2.8)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

 

Low-level exposure

6

1.5 (0.6–3.4)

 

High-level exposure

18

1.8 (1.1–2.9)

 

High probability and high-level exposure

16

1.7 (1.0–2.8)

Cohort Studies—Mortality

Ireland et al., 1997

Benzene-exposed male Monsanto workers

 

 

Production workers

 

 

 

<12 ppm-month

1

1.6 (0.0–8.7)

 

12–72 ppm-month

1

1.9 (0.0–10.5)

 

≥72 ppm-month

2

4.4 (0.5–15.8)

 

Maintenance workers

6

1.2 (0.4–2.5)

Yin et al., 1996a

Chinese factory workers, ever exposed

13

1.3 (0.5–4.1)

Hayes et al., 1996

Benzene-exposed workers in China

 

 

Cumulative exposure

 

 

 

<10 ppm-years

1

0.8

 

10–39 ppm-years

3

1.9

 

40–99 ppm-years

2

1.3

 

100–400 ppm-years

1

0.4

 

400+ ppm-years

5

2.3

 

p trend=0.48

Greenland et al., 1994

White male US transformer-assembly workers, ever exposed, high probability

NA

2.1 (1.0–4.4)

Carpenter et al., 1988

Workers at two nuclear facilities, ever exposed

 

 

No latency

28

0.76

 

10-year-latency

20

0.57

Case-Control Study

Rodvall et al., 1996

Hospital catchment-area residents of Sweden, ever exposed (gliomas)

NA

5.5 (1.4–21.3)

Other Specific Organic Solvents

Cohort Study—Incidence

Anttila et al., 1998

Finnish workers biologically monitored for exposure to aromatic hydrocarbons

 

 

Toluene

3

1.09 (0.22–3.18)

 

Xylene

3

1.62 (0.33–4.72)

Cohort Studies—Mortality

Dosemeci et al., 1991

Male US industrial workers exposed to phenol

10

0.7 (0.4–1.4)

Leffingwell et al., 1983

Male workers at a Texas chemical plant, ever exposed (gliomas)

 

 

Diethylene glycol

11

1.25 (0.48–3.25)b

 

Ethanol

11

1.19 (0.47–3.01)b

 

Isopropanol

13

1.73 (0.59–5.07)b

Case-Control Studies

Rodvall et al., 1996

Hospital catchment-area residents of Sweden, ever exposed (gliomas)

 

 

Toluene

NA

3.4 (0.6–19.3)

 

Xylene

NA

3.3 (0.6–18.6)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Heineman et al., 1994

Male residents of Louisiana, New Jersey, and Philadelphia, ever exposed to chloroform (astrocytic brain tumors)

 

 

All

46

1.0 (0.6–1.6)

 

Low probability

30

0.8 (0.5–1.4)

 

Medium probability

12

3.2 (0.9–12.0)

 

High probability

1

0.2 (0.0–1.8)

Unspecified Mixtures of Organic Solvents

Cohort Study—Incidence

Heineman et al., 1995

Women in Shanghai, China

 

 

Low probability of exposure

89

1.2 (1.0–1.5)

 

High probability of exposure

27

2.0 (1.3–3.0)

 

Low-level exposure

68

1.7 (1.0–1.6)

 

High-level exposure

48

1.5 (1.1–2.0)

 

High probability and high-level exposure

24

1.9 (1.2–2.8)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Mixed solvents, potential routine exposure

13

0.68 (0.36–1.16)

Greenland et al., 1994

White male US transformer-assembly workers, ever exposed

NA

2.65 (0.84–8.36)

Carpenter et al., 1988

Workers at two nuclear facilities, ever exposed

 

 

Trichloroethylene, tetrachloroethylene, or methyl chloroform

 

 

No latency

29

1.76

 

10-year latency

21

1.26

 

Toluene, xylene, or methyl ethyl ketone

 

 

No latency

28

1.96

 

10-year latency

19

1.37

Case-Control Studies

Rodvall et al., 1996

Hospital catchment-area residents of Sweden, ever exposed (gliomas)

 

 

Males

23

2.6 (1.3–5.2)

 

Females

2

0.4 (0.1–2.0)

Heineman et al., 1994

Male residents of Louisiana, New Jersey, and Philadelphia (astrocytic brain tumors)

 

 

Ever exposed

186

1.4 (0.9–1.8)

 

Low probability

48

1.1 (0.6–1.7)

 

Medium probability

32

1.5 (0.8–2.7)

 

High probability

106

1.4 (0.9–2.1)

 

Duration of employment

 

 

2–20 years

80

1.1 (0.7–1.7)

 

21+ years

87

1.7 (1.1–2.6)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Thomas et al., 1987

Male workers in the petroleum-refining and chemical-manufacturing industries (astrocytic brain tumors)

 

 

High exposure to organic solvents

38

1.0

 

<5 years

9

0.9

 

5–19 years

9

0.6

 

≥20 years

14

1.1

NOTE: NA=not available.

aResults from the combined early and recent followup cohort.

b90% CI.

LYMPHATIC AND HEMATOPOIETIC CANCERS

Issue of Classification

Until recently, the International Classification of Diseases (especially ICD-7 and earlier versions) had rubrics for general types of lymphatic and hematopoietic cancers, including lymphosarcoma and reticulosarcoma (ICD-200), Hodgkin’s disease (ICD-201), and lymphatic leukemia (ICD-204). There were no categories for distinguishing specific types of cancers, such as acute leukemia. Thus, in the older epidemiologic studies, all lymphatic and hematopoietic neoplasms were grouped together instead of handled as individual types of cancer (such as Hodgkin’s disease) or specific cell types (such as acute lymphocytic leukemia). The amalgamation of these relatively rare cancers would increase the apparent sample size but could result in diluted estimates of effect if the different sites of cancer were not associated in similar ways with the exposures of interest. In addition, before the use of immunophenotyping to distinguish ambiguous diseases, diagnoses of these cancers may have been misclassified; for example, non-Hodgkin’s lymphoma (NHL) may have been misclassified as Hodgkin’s disease (HD) (Irons, 1992). Misclassification of specific types of cancer, if unrelated to exposure, would have attenuated estimates of relative risk and reduced statistical power to detect associations. When the outcome was mortality, rather than incidence, misclassification would be greater because of the errors in the coding of underlying causes of death on death certificates.

For exposures to organic solvents, the committee reviewed studies that combined all lymphatic and hematopoietic cancers because past exposures were considered substantially higher than the lower permissible exposure limits of today. Given the high exposures of the past, the committee believed that observations of increased risks in some groups of the lymphatic and hematopoietic cancers might be indicative of risk of more specific cancer types. Therefore, the results provided by studies that grouped all the lymphatic and hematopoietic cancers together were used as background information during the committee’s deliberations, and the findings from studies with greater specificity of cancer type were used as the primary evidence in drawing conclusions of associations between exposure to organic solvents and NHL, Hodgkin’s disease (HD), multiple myeloma (MM), acute and adult leukemia, and myelodysplastic syndromes.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

NON-HODGKIN’S LYMPHOMA

In reviewing the literature on NHL, the committee faced an issue of misclassification or selection bias with respect to the accuracy of diagnosis. As explained at the beginning of this section, past ICD codes did not allow for identification of specific cancer types. As a result, epidemiologists may have grouped NHL cases with other lymphatic and hematopoietic cancers to establish sufficient statistical power. The issue was particularly evident when the literature on exposure to nonspecific solvents was reviewed. When a study did not specifically identify NHL as the cancer type being studied, the committee indicated the diagnosis used by the study authors. All these results are presented at the end of this section.

Description of Case-Control Studies

The characteristics of the case-control studies considered by the committee in drawing its conclusions of association are described below in Table 6.33. Most case-control studies reviewed evaluated NHL specifically (Bernard et al., 1984; Blair et al., 1992; Fabbro-Peray et al., 2001; Fritschi and Siemiatycki, 1996b; Hardell et al., 1994; Holly et al., 1997; Olsson and Brandt, 1988; Scherr et al., 1992; Tatham et al., 1997); others studied lymphohematopoietic cancer (Costantini et al. 2001) or malignant lymphoma (Hardell et al., 1981; Persson and Fredriksson, 1999). All case-control studies included interviews with study subjects concerning occupational history, and some interviews included questions about specific chemical exposures.

Some studies relied on the use of a job-exposure matrix (Blair et al., 1992) or industrial hygienist review of questionnaire responses (Costantini et al., 2001; Fritschi and Siemiatycki, 1996b) to determine exposures, but exposure assessment in most of the studies was based on self-reports (Bernard et al., 1984; Fabbro-Peray et al., 2001; Hardell et al., 1981, 1994; Holly et al., 1997; Olsson and Brandt, 1988; Persson and Frederikson, 1999; Persson et al., 1989, 1993; Scherr et al., 1992). Case-control studies on lymphohematopoietic cancer and exposure to organic solvents that had reasonably good assessments of exposure and a sufficient number of exposed cases include those conducted by Blair and colleagues (1992) and Fritschi and Siemiatycki (1996b).

Epidemiologic Studies of Exposure to Organic Solvents and Non-Hodgkin’s Lymphoma

The principal cohort studies that provided information concerning risk of NHL and benzene were the study of Chinese factory workers (Hayes et al., 1997; Yin et al., 1996a,b) and the study of American chemical workers (Wong, 1987a,b). A nested case-control study identified from a cohort of petroleum-distribution workers exposed to benzene at relatively low levels also provided information regarding the association (Schnatter et al., 1996a). Increased relative risks of NHL were found in two analyses in the Chinese study (Hayes et al., 1997: RR=3.0, 95% CI=0.9–10.5; Yin et al., 1996a,b: RR=3.0, 95% CI=1.0–13.0). Risks increased as the duration of occupational exposure to benzene increased, but not as much as with increasing cumulative exposure (Hayes et al., 1997).

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.33 Description of Case-Control Studies of Non-Hodgkin’s Lymphoma and Exposure to Organic Solvents

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Hardell et al., 1981

Male cases, age 25–80 years at diagnosis, admitted to the Department of Oncology in Umea, Sweden and diagnosed in 1974–1978 with histologic confirmation; living controls selected from the National Population Register, matched on sex, age, and place of residence; deceased controls selected from the National Registry for Causes of Death, matched for the above and year of death

105

335

Organic solvents

Self-administered questionnaire with possible telephone followup to assess solvent exposure (self-reports)

Unadjusted

OR

None

Bernard et al., 1984

Cases identified through registries and clinician reporting among residents of the Yorkshire Health Region, UK, with diagnosis in 1979–1981 and histologic confirmation; controls selected from hospital inpatients, matched for age, sex, and geographic area

158

158

Benzene

Solvents

In-person interview to assess occupational history (job titles) and details of solvent and chemical contacts (self-reports)

Maximum likelihood estimate

Sex, age

Olsson and Brandt, 1988

Male cases, age 20–81 years, admitted to the Department of Oncology in Lund, Sweden in 1978–1981, with histologic confirmation; controls selected from previous control groups in studies on Hodgkin’s disease and soft-tissue sarcoma

167

130

Solvents

In-person interview with structured questionnaire to assess lifetime occupational exposure history (self-reports)

Logistic regression

Age, herbicides, chlorophenols

Persson et al., 1989

Cases, age 20–80 years, identified in 1964–1986 at the Orebro Medical Centre Hospital, Sweden; controls randomly selected from population registers (see also Persson and Fredriksson, 1999; Persson et al., 1993)

106

275

Trichloroethylene

White spirit

Solvents

Mailed questionnaire to assess occupational exposures (self-reports)

Unadjusted

OR

None

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Blair et al., 1992

White, male cases identified in Iowa through the Iowa State Health Registry in 1981–1983 and in Minnesota through a surveillance network of hospitals in 1980–1982 with histologic confirmation; controls identified through RDD (alive, age 21–64 years), the records of HCFA (alive, age over 65 years), and vital records (deceased), stratified for state, age, and year of death

622

1,245

Benzene

Solvents

Paints

In-person interview (direct or proxy) with structured questionnaire to assess lifetime occupational history (job titles); probability of exposure determined by industrial hygienist assessment and job-exposure matrixes

Unconditional logistic regression

Age, state, smoking, family cancer history, respondent status, exposure to pesticides and hair dyes

 

Response rates: 87% of cases, 77% of RDD controls, 79% of HCFA controls, 77% of next-of-kin of deceased controls

 

Scherr et al., 1992

Cases identified from nine hospitals in the Boston metropolitan area in 1980–1982 with histologic confirmation; controls selected from town residency lists, matched for sex, age, town, and precinct of residence

303

303

Benzene

Chlorinated solvents

In-person interview or mailed questionnaire to assess lifetime occupational and exposure history (self-reports)

Unadjusted

OR

None

Persson et al., 1993

Cases, age 20–80 years, identified in 1975–1984 at the University Hospital in Linkoping, Sweden, with histologic confirmation; controls randomly selected from population registers (see also Persson and Fredriksson, 1999; Persson et al., 1989)

31

204

Trichloroethylene

White spirit

Solvents

Mailed questionnaire to assess occupational exposures (self-reports)

Logistic regression

Age, other exposures

Hardell et al., 1994

Male cases, age 25–80 years at diagnosis, admitted to the Department of Oncology in Umea, Sweden and diagnosed in 1974–1978 with histologic confirmation; living controls selected from the National Population Register, matched on sex, age, and place of residence; deceased controls selected from the National Registry for Causes of Death, matched for the above and year of death

105

335

Benzene

Trichloroethylene

White spirit

Organic solvents

Self-administered questionnaire with possible telephone followup to assess solvent exposure (self-reports)

Mantel-Haenszel

None

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Fritschi and Siemiatycki, 1996b

Male cases and controls, age 35–70 years, diagnosed in 19 large Montreal-area hospitals in 1979–1985 and histologically confirmed for one of 19 anatomic cancer sites; age-matched, population-based controls were also chosen from electoral lists and RDD

215

1,066 (533 population controls and 533 randomly selected subjects from the eligible cancer control group)

Benzene

Solvents

In-person interviews with segments on work histories (job titles); analyzed and coded by a team of chemists and industrial hygienists (about 300 exposures on semiquantitative scales)

Unconditional logistic regression

Age, proxy status, income, ethnicity

 

Response rates: 83% of cases, 71% of population controls

 

Holly et al., 1997

Male cases identified through the Northern California Cancer Center from hospitals in six counties; controls selected through RDD, and matched for sex, county, and age

312

420

Benzene

Chlorinated solvents

In-person interview with structured questionnaire to assess chemical and occupational exposures (self-reports)

Unconditional logistic regression

Age

Tatham et al., 1997

Male cases from the Selected Cancers Study, born in 1929–1953, identified from eight population-based cancer registries in 1984–1988 with histologic confirmation; controls selected through RDD and matched for registry and date of birth

1,048

1,659

Chlorinated hydrocarbons

Chemical solvents

Telephone interviews (direct or proxy) with questionnaire to assess work (job titles) and exposure history (self-reports)

Conditional logistic regression

Age at diagnosis, year entering study, ethnicity, education, Jewish religion, marital status, smoking status, service in Vietnam, other medical factors

 

Response rates: 88% of cases, 83% of controls

 

Persson and Fredriksson, 1999

Cases, age 20–80 years, identified in 1964–1986 at the Orebro Medical Centre Hospital and in 1975–1984 at the University Hospital in Linkoping, Sweden, with histologic confirmation; controls randomly selected from population registers (see also Persson et al., 1989, 1993)

199

479

Benzene

Trichloroethylene

White spirit

Solvents

Mailed questionnaire to assess occupational exposures (self-reports)

Mantel-Haenszel

Age, sex

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Costantini et al., 2001

Cases, age 20–74 years, identified through periodic hospital survey and diagnosed in 12 regions in Italy in 1991–1993 with histologic confirmation; controls randomly selected from municipal demographic files and the National Health Services files, matched for age and sex

1,450

1,779

Launderers, dry cleaners, pressers

Painters

Printers

In-person interview (direct or proxy) with standardized and job-specific questionnaires to assess lifetime occupational history and exposure to solvents; probability of exposure further determined by industrial hygienist

Mantel-Haenszel

Age, sex

 

Response rates: 88% of cases, 81% of controls

 

Fabbro-Peray et al., 2001

Cases, age 18 years or over, identified from hospitals in Languedoc-Roussillon, France and diagnosed in 1992–1995 with histologic confirmation; controls randomly selected from electoral lists

445

1,025

Benzene

Dry-cleaning solvents

Rubber industry

Paints

In-person interviews to assess occupational history (job titles) and exposure to specific chemicals (self-reports)

Unconditional logistic regression

Age, sex, education, urban setting

NOTE: HCFA=Health Care Financing Administration; RDD=random-digit dialing.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

The US chemical-worker study (Wong, 1987a,b) had only five exposed cases. It showed a weak association with continuous exposure (SMR=1.13, 95% CI=0.23–3.30) and no increase in risk with increasing cumulative exposure. The authors examined duration of exposure categorized, but there was only one exposed case per category. In the study of petroleum distributors, with eight exposed cases (Schnatter et al., 1996a), relative risks did not increase with increasing levels of exposure. Wilcosky and colleagues (1984) found a 3-fold risk among rubber-plant workers (OR=3.0; the authors did not provide CIs, and the committee was not able to calculate them), and Bond and colleagues (1986) in a study of chemical workers found about a 100% increase in mortality (MSR=1.99, 95% CI=0.24–7.22). Another cohort study of benzene-exposed oil-refinery workers (Tsai et al., 1983) reported no exposed cases.

Among the case-control studies in which the association between exposure to benzene and NHL was assessed, increased risks were found in the studies by Fabbro-Peray and colleagues (2001) (OR=2.0, 95% CI=1.1–3.9); odds ratios increased with duration of exposure. Blair and colleagues (1992) found increased odds ratios for “higher” exposures (OR=1.5, 95% CI=0.7–3.1), and Hardell and colleagues (1994) found very large excess risks associated with self-reported exposure to benzene (OR=28, 95% CI=1.8–730). No associations were found in several other small case-control studies (Bernard et al., 1984; Fritschi and Siemiatycki, 1996b; Holly et al., 1997; Persson and Fredriksson, 1999; Scherr et al., 1992; Schumacher and Delzell, 1988).

The cohort studies of trichloroethylene-exposed workers and NHL risk include studies of Finnish and Danish biologically monitored workers (Anttila et al., 1995; Hansen et al., 2001), US aircraft and aerospace maintenance and manufacturing workers (Blair et al., 1998, Boice et al., 1999; Morgan et al., 1998), and workers in Sweden (Axelson et al., 1994).

In the Finnish biologic-monitoring study (Anttila et al., 1995), relative risks were increased (SIR=2.01, 95% CI=0.65–4.69). In the Danish cohort study (Hansen et al., 2001) of workers biologically monitored for trichloroethylene, a 3.5-fold increased risk of NHL was observed (95% CI=1.5–6.9). The risk did not increase with increasing levels of exposure, but the statistical power was too low to detect trends.

Increased NHL mortality was found in the US aircraft-maintenance worker study in Utah (Blair et al., 1998) (SMR=2.0, 95% CI=0.9–4.6), but it did not increase with cumulative exposure among male and female workers. A 140% increase in risk (OR=2.4; CI was not provided in the study, and the committee was not able to calculate it) was also found in rubber workers exposed to trichloroethylene (Wilcosky et al., 1984). In the study of aircraft-manufacturing workers in California (Boice et al., 1999), no overall increase in risk was found (SMR=1.19, 95% CI=0.65–1.99), but relative risks increased as duration of exposure increased. Axelson and colleagues (1994) found an 85% increase in incidence (SIR=1.85, 95% CI=0.38–5.41) among Swedish workers, but no association was found (SMR=0.96, 95% CI=0.20–2.81) in the study of aerospace workers in Arizona (Morgan et al., 1998). Two case-control studies in Sweden found increases in NHL risk of 7.2 (95% CI=1.3–42) (Hardell et al., 1994) and 1.2 (95% CI=0.5–2.4) (Persson and Fredriksson, 1999) for NHL, but both were based on self-reported exposures.

In several cohort and case-control studies, relative risks of NHL were calculated according to exposure to tetrachloroethylene and dry-cleaning solvents (Anttila et al., 1995; Blair et al., 1990; Boice et al., 1999; Costantini et al., 2001; Fabbro-Peray et al., 2001), toluene or xylene (Anttila et al., 1998; Blair et al., 1998; Svensson et al., 1990; Wilcosky et al., 1984), phenol (Dosemeci et al., 1991), white spirits (Persson and Fredrickson, 1999; Hardell et al.,

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

1994), and other specific solvents (Blair et al, 1998; Wilcosky et al., 1984). No consistent positive associations were found between NHL and any of those solvents except white spirits, with which positive associations were found by Persson and Fredrickson (1999) in a Swedish study (OR=2.6, 95% CI=1.3–4.7) and by Hardell and colleagues (1994) (OR=3.2, 95% CI=1.3–8.3).

The association between exposure to mixtures of organic solvents and NHL has been examined in several studies, many of which are occupational studies, including aircraft manufacturing and maintenance workers (Blair et al., 1998; Boice et al., 1999), uranium-processing plant workers (Ritz, 1999), painters (Blair et al., 1992; Costantini et al., 2001; Lundberg and Milatou-Smith, 1998; Steenland and Palu, 1999), printers (e.g., Costantini et al., 2001; Leon, 1994; Nielsen et al., 1996), and people in several other occupations (such as shoe manufacturing and rubber work). Some studies showed positive associations, and others showed associations close to unity.

Summary and Conclusion

Although there is a substantial body of literature on the association between exposure to specific organic solvents and solvent mixtures and risk of NHL, most studies are based on small numbers of exposed cases. An association between comparatively high relative risks of NHL and exposure to benzene was seen consistently in a number of cohort studies. The studies on exposure to benzene and NHL provide consistently positive findings because the populations or groups had known exposure and there was evidence of exposure-response relationships.

The committee concludes, from its assessment of the epidemiologic literature, that there is limited/suggestive evidence of an association between chronic exposure to benzene and non-Hodgkin’s lymphoma.

Two cohort studies suggested that there was an increased risk of dying from NHL with exposure to trichloroethylene (Blair et al., 1998; Wilcosky et al., 1984). However, in the rubber worker study, (Wilcosky et al., 1984) they were exposed to numerous other chemicals in addition to trichloroethylene. The Blair study did not demonstrate an exposure-response relationship in examining mortality risk and found no association in relation to incidence risk. One case-control study showed a very strong association (Hardell et al., 1994), but the committee did not feel that it constituted compelling evidence inasmuch there was a high probability that the relative risks were overstated. The evidence between exposure to white spirit and other specific organic solvents was also limited by Hardell and colleagues’ study (1994) that used self-reported exposures and lacked credibility owing to possible recall bias. For exposure to unspecified mixtures of solvents, as is the case with many of the other exposures discussed above, statistical fluctuations made it difficult to form any conclusion. Table 6.34 identifies the studies reviewed by the committee in drawing its conclusion regarding association. Unless indicated in the table, the study populations include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to the solvents under review, other than benzene, and non-Hodgkin’s lymphoma.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.34 Selected Epidemiologic Studies—Non-Hodgkin’s Lymphoma and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Studies—Incidence

Hansen et al., 2001

Biologically monitored Danish workers

 

 

Male workers, ever exposed

8

3.5 (1.5–6.9)

 

Cumulative exposure:

 

 

Unknown

2

3.6 (0.4–13.0)

 

<1080 months, mg/m3

3

3.9 (0.8–11.0)

 

≥1080 months, mg/m3

3

3.1 (0.6–9.1)

Blair et al, 1998

Aircraft-maintenance workers in Utah

 

 

Cumulative exposure, males (incidence)

 

 

No exposure

5

0.5 (0.2–1.7)

 

<5 unit-years

8

0.9 (0.3–2.6)

 

5–25 unit-years

4

0.7 (0.2–2.6)

 

>25 unit-years

7

1.0 (0.4–2.9)

 

Cumulative exposure, females (incidence)

 

 

No exposure

0

 

<5 unit-years

1

0.6 (0.1–5.0)

 

5–25 unit-years

0

 

>25 unit-years

2

0.9 (0.2–4.5)

Anttila et al., 1995

Finnish workers exposed to halogenated hydrocarbons

8

1.81 (0.78–3.56)

 

Exposure level

 

 

<100 µmol/L

5

2.01 (0.65–4.69)

 

100 + µmol/L

2

1.40 (0.17–5.04)

Axelson et al., 1994

Swedish workers

5

1.56 (0.51–3.64)

 

≥2 years of exposure and 10-year latency

3

1.85 (0.38–5.41)

 

49 mg/L

2

1.64 (0.20–5.92)

 

50–99 mg/L

0

 

100 + mg/L

1

8.33 (0.22–46.43)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Trichloroethylene, ever exposed (routine)

14

1.19 (0.65–1.99)

 

Years of exposure (routine or intermittent)

 

 

<1

7

0.74 (0.32–1.72)

 

1–4

10

1.33 (0.64–2.78)

 

≥5

14

1.62 (0.82–3.22)

Blair et al., 1998

Aircraft-maintenance workers in Utah, ever exposed

28

2.0 (0.9–4.6)

 

Cumulative exposure, males (mortality)

 

 

No exposure

11

1.6 (0.5–4.5)

 

<5 unit-years

10

1.8 (0.6–5.4)

 

5–25 unit-years

6

1.9 (0.6–6.3)

 

>25 unit-years

5

1.1 (0.3–3.8)

 

Cumulative exposure, females (mortality)

 

 

No exposure

2

2.0 (0.3–12.2)

 

<5 unit-years

3

3.8 (0.8–18.9)

 

5–25 unit-years

0

 

>25 unit-years

4

3.6 (0.8–16.2)

 

Exposure levels (males)

 

 

Low-level intermittent

15

1.5 (0.5–4.3)

 

Low-level continuous

12

1.8 (0.6–5.2)

 

Frequent peaks

9

1.5 (0.5–4.4)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

 

Exposure levels (females)

 

 

Low-level intermittent

4

3.9 (0.8–17.7)

 

Low-level continuous

2

3.4 (0.5–21.7)

 

Frequent peaks

5

3.8 (0.9–16.2)

Morgan et al., 1998

Aerospace workers in Arizona

 

 

Trichloroethylene-exposed subcohort

 

 

Lymphosarcoma and reticulosarcoma

3

0.96 (0.20–2.81)

 

All other lymphopoietic tissue

11

1.01 (0.51–1.81)

Wilcosky et al., 1984

Male US rubber-industry workers, exposed >1 year

3

2.4

Case-Control Studies

Persson and Fredriksson, 1999

Residents of Sweden

 

1+ years exposed, latency 5–45 years

16

1.2 (0.5–2.4)

Hardell et al., 1994

Male residents of Sweden

 

 

Ever exposed

4

7.2 (1.3–42)

Benzene

Cohort Studies—Incidence

Hayes et al., 1997

Chinese factory workers

16

3.0 (0.9–10.5)

 

Exposed workers, year of hire

 

 

<1972

15

4.1

 

≥1972

1

0.5

 

Exposed workers, average ppm

 

 

<10

7

2.7 (0.7–10.6)

 

10–24

2

1.7 (0.3–10.2)

 

≥25

7

4.7 (1.2–18.1)

 

Exposed workers, duration

 

 

<5 years

1

0.7 (0.1–7.2)

 

5–9 years

4

3.3 (0.7–14.7)

 

≥10 years

11

4.2 (1.1–15.9)

 

Exposed workers, cumulative exposure

 

 

<40 ppm-years

6

3.3 (0.8–13.1)

 

40–99 ppm-years

1

1.1 (0.1–11.1)

 

≥100 ppm-years

9

3.5 (0.9–13.2)

Yin et al., 1996a,b

Chinese factory workers, ever exposed

17

3.0 (1.0–13.0)

Cohort Studies—Mortality

Schnatter et al., 1996a

Male Canadian petroleum-distribution workers

 

 

0.00–0.49 ppm-years

4

1.00

 

0.50–7.99 ppm-years

3

1.21 (0.16–8.07)

 

8.00–19.99 ppm-years

1

1.14 (0.02–22.1)

 

20–219.8 ppm-years

0

0.0 (0.0–27.6)

Wong, 1987a,b

Male US chemical workers

5

0.91 (0.29–2.11)

 

Continuously exposed

3

1.13 (0.23–3.30)

 

Continuous exposure by duration

 

 

<5 years

1

0.65

 

5–14 years

1

1.56

 

≥15 years

1

2.18

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

 

Cumulative exposure

 

 

<180 ppm-months

3

2.57 (0.24–3.41)

 

180–719 ppm-months

3

1.61 (0.38–5.45)

 

≥720 ppm-months

1

1.34 (0.19–4.14)

 

p-trend=0.71

Bond et al., 1986

Male US chemical workers

 

 

Lymphosarcoma and reticulosarcoma

 

 

Entire cohort

2

1.80 (0.22–6.57)

 

Cohort less those exposed to arsenic, asbestos, or high levels of vinyl chloride

2

1.99 (0.24–7.22)

 

Other lymphatic tissue

 

 

Entire cohort

1

0.76 (0.02–4.29)

 

Cohort less those exposed to arsenic, asbestos, or high levels of vinyl chloride

1

0.84 (0.02–4.64)

Wilcosky et al., 1984

Male US rubber-industry workers, exposed >1 year

6

3.0

Tsai et al., 1983

Texas refinery workers

 

 

1+ years employment in benzene areas

 

 

All lymphopoietic cancer

0

Case-Control Studies

Fabbro-Peray et al., 2001

Residents of France, ever exposed

22

2.0 (1.1–3.9)

≤15 years exposed

9

1.7 (0.7–4.3)

 

>15 years exposed

13

2.4 (0.9–5.9)

Persson and Fredriksson, 1999

Residents of Sweden

 

1+ years exposed, latency 5–45 years

3

0.8 (0.1–3.8)

Holly et al., 1997

Male residents of San Francisco Bay Area

 

 

<10 hours of exposure

294

1.0

 

10+ hours of exposure

17

1.2 (0.62–2.4)

Fritschi and Siemiatycki, 1996b

Male residents of Montreal

 

Nonsubstantial

20

0.7 (0.4–1.1)

 

Substantial

6

0.8 (0.3–2.1)

Hardell et al., 1994

Male residents of Sweden, ever exposed

3

28 (1.8–730)

Blair et al., 1992

Male residents of Iowa and Minnesota

 

 

Potential exposure

153

1.1 (0.9–1.4)

 

Lower intensity

141

1.1 (0.8–1.4)

 

Higher intensity

12

1.5 (0.7–3.1)

Scherr et al., 1992

Residents of Boston, MA, ever exposed

4

1.2 (0.5–2.6)

Schumacher and Delzell, 1988

Residents of North Carolina

 

Occupational exposure (ever):

 

 

White males

56

0.77 (0.56–1.07)

 

Black males

10

0.94 (0.47–1.87)

Bernard et al., 1984

Residents of Yorkshire, England, benzene use

 

 

Males

NA

0.49 (0.21–2.00)

Other Specific Organic Solvents

Cohort Studies—Incidence

Anttila et al., 1998

Finnish workers monitored for exposure to solvents

 

 

Toluene or xylene, all years

3

1.18 (0.24–3.45)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Anttila et al., 1995

Finnish workers biologically monitored for halogenated hydrocarbons

 

 

Tetrachloroethylene

3

3.76 (0.77–11.0)

 

1,1,1-Trichloroethane

1

3.87 (0.10–21.5)

Svensson et al., 1990

Male rotogravure-plant workers in Sweden

 

 

Toluene

 

 

All exposed

1

0.33 (0.01–1.86)

 

≥5 years of exposure, >10-year latency

1

0.52 (0.01–2.91)

Cohort Studies-Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Tetrachloroethylene, ever exposed (routine)

8

1.70 (0.73–3.34)

 

Years of exposure (routine or intermittent)

 

 

<1

4

1.25 (0.43–3.57)

 

1–4

6

1.11 (0.46–2.70)

 

≥5

10

1.41 (0.67–3.00)

Blair et al., 1998

Aircraft maintenance workers in Utah

 

 

Males, ever exposed

 

 

Stoddard solvent

16

1.3 (0.5–3.5)

 

Isopropyl alcohol

7

1.8 (0.6–5.8)

 

Other alcohols

3

2.1 (0.5–9.0)

 

Acetone

6

1.7 (0.5–5.5)

 

Toluene

3

1.0 (0.2–4.2)

 

Methyl ethyl ketone

4

1.4 (0.4–5.1)

 

Methylene chloride

6

3.0 (0.9–10.0)

 

Females, ever exposed

 

 

Stoddard solvent

5

2.4 (0.6–9.9)

 

Isopropyl alcohol

2

5.8 (1.0–34.6)

 

Other alcohols

0

 

Acetone

1

1.3 (0.1–12.8)

 

Toluene

2

2.2 (0.4–13.1)

 

Methyl ethyl ketone

1

1.6 (0.2–15.7)

 

Methylene chloride

0

Dosemeci et al., 1991

White, male chemical-manufacturing workers

 

 

Phenol, ever exposed

4

0.4 (0.1–1.1)

Blair et al., 1990

Members of a dry-cleaning union in St. Louis, MO

 

 

Lymphosarcoma or reticulosarcoma

7

1.7 (0.7–3.4)

 

Other lymphatic cancers

4

0.7 (0.2–1.8)

Wilcosky et al., 1984

US rubber-industry workers, exposed >1 year

 

 

Specialty naphthas

6

1.4

 

Xylenes

4

3.7

 

Isopropanol

6

2.9

Case-Control Studies

Costantini et al., 2001

Residents of 12 areas of Italy

 

 

Launderers, dry cleaners, and pressers

3

1.6 (0.3–9.1)

Fabbro-Peray et al., 2001

Residents of France

 

Dry-cleaning solvents, ever exposed

35

1.0 (0.6–1.6)

Persson and Fredriksson, 1999

Residents of Sweden

 

White spirit, 1+ years exposed, latency 5–45 years

27

2.6 (1.3–4.7)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Hardell et al., 1994

Male residents of Sweden

 

 

White spirit (ever exposed)

12

3.2 (1.3–8.3)

Unspecified Mixtures of Organic Solvents

Cohort Studies—Incidence

Anttila et al., 1998

Biologically monitored Finnish workers

 

 

Aromatic hydrocarbons, all years

2

0.52 (0.06–1.88)

Lundberg and Milatou-Smith, 1998

Swedish paint-industry workers

 

Male workers, ≤5 years of employment

2

1.0 (0.1–3.4)

Nielsen et al., 1996

Danish lithographer-union members, malignant lymphoma

2

1.5 (0.3–5.0)

Berlin et al., 1995

Swedish workers occupationally exposed to solvents

7

1.9 (0.8–4.0)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

All factory workers

137

0.94 (0.79–1.11)a

 

Mixed solvents, ever exposed (routine)

29

1.02 (0.68–1.47)

 

Years of exposure (routine or intermittent)

 

 

<1

9

0.77 (0.37–1.62)

 

1–4

20

0.79 (0.45–1.39)

 

≥5

52

1.01 (0.64–1.61)

Ritz, 1999

White male US uranium-processing plant workers, employed >3 months

8

1.71 (0.73–3.36)

Steenland and Palu, 1999

US painter union members

 

Painters (all members)

137

1.06 (0.89–1.25)

 

Painters (20 years since first union membership)

110

1.10 (0.91–1.32)

Blair et al., 1998

Aircraft-maintenance workers in Utah

 

 

Any solvent, ever exposed

 

 

Males

31

1.6 (0.6–4.1)

 

Females

9

2.8 (0.8–4.5)

 

Solvents, unspecified, ever exposed

 

 

Males

31

1.6 (0.6–4.1)

 

Females

9

2.9 (0.8–10.6)

Fu et al., 1996

Shoe-manufacturing workers

 

 

English cohort

6

0.55 (0.20–1.20)

 

Probable solvent exposure

2

0.54 (0.07–1.96)

 

High solvent exposure

0

 

Florence cohort

2

1.06 (0.13–3.82)

 

Probable solvent exposure

2

2.44 (0.30–8.81)

 

High solvent exposure

2

2.74 (0.33–9.90)

Hunting et al., 1995

Male vehicle mechanics in Washington, DC

 

 

Solvents or fuels, high exposure (lymphatic and hematopoietic)

3

4.22 (0.87–12.34)

 

Solvents or fuels, high exposure (other neoplasms of lymphatic and hematopoietic)

1

2.57 (0.06–14.27)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Leon, 1994

Male trade-union members in the British printing industry

 

 

Machine managers

2

1.21 (0.15–4.37)

 

Compositors

4

1.17 (0.32–2.99)

 

Machine assistants

4

2.16 (0.59–5.54)

 

Publishing-room men

2

2.68 (0.32–9.68)

Wong et al., 1993

US gasoline-distribution workers

 

 

Land-based terminal cohort (cancer of other lymphatic tissue)

18

0.92 (0.54–1.45)

 

Marine-based cohort (cancer of other lymphatic tissue)

11

0.73 (0.39–1.25)

Walker et al., 1993

Shoe-manufacturing workers in Ohio, employed >1 month

5

0.75 (0.24–1.77)

Teta et al., 1992

Male workers at ethanol and isopropanol production plants in Texas

 

 

All workers (South Charlestown plant)

5

5.60 (1.8–13.0)

Paci et al., 1989

Shoe-manufacturing plant in Florence, Italy, ever employed

 

 

Other lymphatic and hematopoietic neoplasms

 

 

Males

1

0.55 (0.0–3.1)a

 

Females

1

1.11 (0.0–6.2)a

Sorahan et al., 1989

British rubber-industry workers (lymphoid cancers)

50

0.91 (0.68–1.21)a

Garabrant et al., 1988

Aircraft-manufacturing workers in California, employed >4 years

 

 

Lymphosarcoma and reticulosarcoma

13

0.82 (0.44–1.41)

 

Other neoplasms of lymphatic and hematopoetic tissue

5

0.65 (0.21–1.52)

Wilcosky et al., 1984

Male US rubber-industry workers, exposed >1 year

 

 

Solvent “A” (mixture of toluene and other solvents)

6

2.6

Paganini-Hill et al., 1980

Newspaper web pressmen union members in Los Angeles, CA (other lymphatic and hematopoietic neoplasms)

5

1.29 (0.42–2.99)a

Case-Control Studies

Costantini et al., 2001

Residents of 12 areas of Italy

 

 

Painters, ever employed

20

1.2 (0.6–2.4)

 

Printers, ever employed

11

1.2 (0.6–2.9)

Fabbro-Peray et al., 2001

Residents of France

 

Rubber industry, ever employed

16

1.6 (0.8–3.4)

 

Paints, ever exposed

26

0.8 (0.5–1.3)

Persson and Fredriksson, 1999

Residents of Sweden

 

1+ years exposed, latency 5–45 years

 

 

Solvents

61

1.6 (1.0–2.5)

 

Solvents, high intensity

51

1.8 (1.1–2.9)

Holly et al., 1997

Male residents of the San Francisco Bay area

 

 

Chlorinated solvents

 

 

<10 hours of exposure

291

1.0

 

10+ hours of exposure

20

0.73 (0.41–1.3)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Tatham et al., 1997

Males in US cancer registries

 

 

Chemical solvents, ever exposed

451

1.10 (0.90–1.30)

 

Chlorinated hydrocarbons, ever exposed

49

0.91 (0.62–1.30)

Fritschi and Siemiatycki, 1996b

Male residents of Sweden

 

Nonsubstantial

26

0.6 (0.4–1.0)

 

Substantial

48

0.9 (0.6–1.3)

Hardell et al., 1994

Male residents of Sweden

 

 

Organic solvents

 

 

All

45

2.4 (1.4–3.9)

 

High grade

31

2.9 (1.6–5.6)

 

Low grade

14

1.8 (0.8–3.8)

Partanen et al., 1993

Finnish production workers in the wood industry (leukemias and lymphomas pooled)

 

 

Solvents, ever exposed

4

5.62 (0.99–32.0)

 

Solvents, adjusted for formaldehyde

4

5.07 (0.40–63.5)

Blair et al., 1992

Male residents of Iowa and Minnesota

 

 

Solvents (excluding benzene)

 

 

Potentially exposed

359

1.1 (0.9–1.4)

 

Lower intensity

334

1.1 (0.8–1.4)

 

Higher intensity

25

1.4 (0.8–2.5)

 

Paints:

 

 

Potentially exposed

116

1.1 (0.9–1.5)

 

Lower intensity

107

1.1 (0.9–1.5)

 

Higher intensity

9

1.1 (0.5–2.6)

Scherr et al., 1992

Residents of Boston, MA

 

 

Chlorinated solvents, ever exposed

24

1.2 (0.8–1.8)

Persson et al., 1989

Residents of Sweden

 

 

Solvents

33

2.0 (1.2–3.5)a

 

Solvents, high intensity

27

2.1 (1.2–3.7)a

Olsson and Brandt, 1988

Male residents of Sweden

 

Solvent exposure, ever (based on observed cases)

NA

2.0 (1.5–2.6)

 

120 months

NA

1.8 (1.2–2.7)

 

240 months

NA

3.3 (1.5–7.1)

 

360 months

NA

6.0 (1.9–19.0)

Bernard et al., 1984

Residents of Yorkshire, England

 

 

Occupational solvents (excluding benzene)

 

 

Males

NA

1.52 (0.70–3.26)

 

Females

NA

7.71 (1.24–47.93)

 

Solvent use as a hobby, males

NA

1.39 (0.30–6.46)

Hardell et al., 1981

Male residents of Sweden (malignant lymphoma)

 

 

Organic solvents, low grade

10

1.2 (0.5–2.6)

 

Organic solvents, low and high grade

50

2.4 (1.5–3.8)

 

Solvents, phenoxy acids, or chlorophenols

23

8.5 (4.2–17.2)

NOTE: NA=not available.

a95% CI was calculated by the committee with standard methods from the observed and expected numbers presented in the original papers.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

HODGKIN’S DISEASE

Description of Case-Control Studies

The characteristics of the case-control studies considered by the committee in drawing its conclusion regarding association are described in Table 6.35. All case-control studies of HD reviewed by the committee included interviews with study subjects about occupational history; some also inquired about specific chemical exposures. Although some studies relied on industrial hygienist review of questionnaire responses (Costantini et al., 2001) to determine exposure, exposure assessment in other studies was based on self-reports (Bernard et al., 1984; Hardell and Bengtsson, 1983; Persson et al., 1989, 1993).

Epidemiologic Studies Regarding Exposure to Organic Solvents and Hodgkin’s Disease

Each of the studies of exposure to trichloroethylene reviewed showed an increased risk of HD. However, most studies had small numbers of exposed cases (Anttila et al., 1995; Axelson et al., 1994; Blair et al., 1998; Boice et al., 1999; Morgan et al., 1998; Persson et al., 1993); therefore, the confidence intervals were broad and included unity. The only study showing markedly increased risk was by Persson and colleagues (1989) (OR=2.8, 95% CI 0.96–7.86).

Two studies provided results on the relationship between exposure to benzene and risk of HD. A cohort study of male chemical-plant workers (Wong, 1987a) who were continuously exposed to benzene did not show any increased risk (SMR=1.12, 95% CI=0.14–4.05), and a case-control study among residents in Yorkshire, England (Bernard et al., 1984) also showed no relationship between exposure to benzene and HD (OR=1.00, 95% CI=0.50–1.50).

Studies of other specific solvents—including tetrachloroethylene (Blair et al., 1990; Boice et al., 1999), toluene and xylene (Anttila et al., 1998), phenol (Dosemeci et al., 1991), and white spirit (Persson et al., 1989, 1993)—did not show any evidence of increased risk of HD.

The risk of HD associated with exposure to unspecified mixtures of solvents or employment in solvent-related occupations—including aircraft-manufacturing workers, painters, printers, and dry cleaners—was investigated in several cohort and case-control studies. Although several studies yielded positive relative risks, most had considerable statistical variability in their estimates. Only one study (Hardell and Bengston, 1983) showed a risk that was well above the null value; its subjects had concomitant exposure to phenoxy acids and chlorophenols, and exposure to all substances conferred an almost 7-fold excess risk (OR=6.6, 95% CI 2.4–18.5).

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.35 Description of Case-Control Studies of Hodgkin’s Disease and Exposure to Organic Solvents

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Hardell and Bengtsson, 1983

Male cases, age 25–80 years at diagnosis, admitted to the Department of Oncology in Umea, Sweden, and diagnosed in 1974–1978 with histologic confirmation; living controls selected from the National Population Register, matched on sex, age, and place of residence; deceased controls selected from the National Registry for Causes of Death, matched for the above and year of death

60

335

Organic solvents

Self-administered questionnaire with possible telephone followup to assess solvent exposure (self-reports)

Mantel-Haenszel

None

Bernard et al., 1984

Cases identified through registries and clinician reporting among residents of the Yorkshire Health Region, UK, with diagnosis in 1979–1981 and histologic confirmation; controls selected from hospital inpatients, matched for age, sex, and geographic area

48

48

Benzene

Solvents

In-person interview to assess occupational history (job titles) and details of solvent and chemical contacts (self-reports)

Logistic regression

Sex, age

Persson et al., 1989

Cases, age 20–80 years, identified in 1964–1986 at the Orebro Medical Centre Hospital, Sweden; controls randomly selected from population registers

54

275

Trichloroethylene

White spirit

Solvents

Mailed questionnaire to assess occupational exposures (self-reports)

Unadjusted OR,

Logistic regression

Age, sex, other exposures

Persson et al., 1993

Cases, age 20–80 years, identified in 1975–1984 at the University Hospital in Linkoping, Sweden, with histologic confirmation; controls randomly selected from population registers

31

204

Trichloroethylene

White spirit

Solvents

Mailed questionnaire to assess occupational exposures (self-reports)

Unadjusted OR,

Logistic regression

Age, other exposures

Costantini et al., 2001

Cases, age 20–74 years, identified through periodic hospital survey and diagnosed in 12 regions in Italy in 1991–1993 with histologic confirmation; controls randomly selected from municipal demographic files and the National Health Services files, matched for age and sex

365

1,779

Launderers, dry cleaners, pressers

In-person interview (direct or proxy) with standardized and job-specific questionnaires to assess lifetime occupational history (job titles) and exposure to solvents (self reports); probability of exposure further determined by industrial hygienist

Mantel-Haenszel

Age, sex

 

Response rates: 88% of cases, 81% of controls

 

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Summary and Conclusion

Overall, the studies reviewed by the committee did not show any persuasive evidence of associations between HD and exposure to specific solvents or solvent mixtures. Although many of the studies of exposure to mixtures of solvents yielded increased risk estimates, there was considerable statistical variability in them. The incidence of HD is low, and most studies had small numbers of exposed cases to evaluate. Having such small numbers may lead to spuriously increased relative risks when the null hypothesis of no association is true. That limitation is reflected in the wide CIs observed in most of the studies. The lack of specific or validated exposure-assessment information and the impact of bias are other limitations that the committee considered in drawing its conclusion. Table 6.36 identifies the key studies reviewed for each exposure and the data points evaluated by the committee. Unless indicated in the tables, the study populations include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to solvents under review and Hodgkin’s disease.

TABLE 6.36 Selected Epidemiologic Studies—Hodgkin’s Disease and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Study—Incidence

Anttila et al., 1995

Biologically monitored workers in Finland

3

1.70 (0.35–4.96)

 

<100 µmol/L

2

2.00 (0.24–7.22)

 

≥100 µmol/L

1

1.83 (0.05–10.2)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California, potential routine exposure

4

2.77 (0.76–7.10)

Blair et al., 1998

Aircraft-maintenance workers in Utah, employed >1 year

5

1.4 (0.2–12.0)

Morgan et al., 1998

Aerospace workers in Arizona

 

 

All exposed

1

0.60 (0.02–3.35)

 

Low exposure

1

1.55 (0.04–8.64)

 

High exposure

0

Axelson et al., 1994

Biologically monitored male workers in Sweden

1

1.07 (0.03–5.95)

Case-Control Study

Persson et al., 1989

Residents of Sweden, exposed >1 year

7

2.8 (0.96–7.86)a

Benzene

Cohort Study—Mortality

Wong, 1987a,b

US male chemical-plant workers

3

0.81 (0.16–2.36)

 

Benzene, continuously exposed

2

1.12 (0.14–4.05)

 

Not exposed to benzene

1

0.75 (0.19–4.15)

 

Duration of exposure, continuously exposed to benzene:

 

 

<5 years

1

0.88

 

5–14 years

1

2.37

 

≥15 years

0

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Case-Control Study

Bernard et al., 1984

Residents of Yorkshire, England

NA

1.00 (0.50–1.50)

Other Specific Organic Solvents

Cohort Study—Incidence

Anttila et al., 1998

Biologically monitored workers in Finland (includes HD, NHL, and other lymphohematopoietic cancers)

3

0.78 (0.16–2.28)

 

Toluene

3

1.18 (0.24–3.45)

 

Xylene

 

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Tetrachloroethylene, potential routine exposure

0

Dosemeci et al., 1991

US male plant workers

 

 

Phenol

 

 

Any exposure

10

1.7 (0.8–3.1)

 

No exposure

2

0.5

 

Low cumulative exposure

8

2.3

 

Medium cumulative exposure

2

0.9

 

High cumulative exposure

0

Blair et al., 1990

Members of a dry-cleaning union in Missouri

 

 

Dry-cleaning solvents

4

2.1 (0.6–5.3)

Case-Control Studies

Persson et al., 1993

Residents of Sweden

 

 

White spirits, exposed >1 year

4

1.4 (0.36–4.67)a

Unspecified Mixtures of Organic Solvents

Cohort Studies—Incidence

Anttila et al., 1998

Biologically monitored workers in Finland

 

 

Aromatic hydrocarbons

3

1.49 (0.31–4.36)

Anttila et al., 1995

Biologically monitored workers in Finland

 

 

Halogenated hydrocarbons

3

1.45 (0.30–4.23)

Berlin et al., 1995

Swedish workers occupationally exposed to solvents

2

6.3 (0.8–22.7)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Mixed solvents, potential routine exposure

6

1.61 (0.59–3.51)

Ritz, 1999

White male uranium-processing workers in Ohio

 

 

Ever exposed to solvents

6

2.09 (0.76–4.54)

Steenland and Palu, 1999

Members of the US Painters Union

 

 

Ever employed as a painter

16

1.30 (0.74–2.11)

 

Nonpainter

4

0.54 (0.06–1.93)

 

20 years since first union membership

10

1.17 (0.56–2.15)

Walker et al., 1993

Ohio shoe-manufacturing plants, employed >1 month

4

1.12 (0.31–2.88)

Garabrant et al., 1988

Aircraft-manufacturing workers in California, employed >4 years

4

0.73 (0.20–1.88)

Wen et al., 1985

Cohort of male oil refinery workers

 

 

Total lubricating cohort

1

1.61 (0.04–8.98)

 

Other lubricating cohort

1

2.17 (0.28–12.10)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Alderson and Rattan, 1980

Male British workers in dewaxing plants

 

Isopropyl alcohol production-plant workers

0

0.09 (0.0–33.55)b

 

Methyl ethyl ketone production-plant workers

0

0.10 (0.0–20.50)b

Case-Control Studies

Costantini et al., 2001

Residents in 12 areas of Italy

 

 

Launderers, dry cleaners, and pressers

1

2.5 (0.3–24.6)

Persson et al., 1993

Residents of Sweden

 

 

Solvents, high intensity

7

1.8 (0.7–4.6)c

Bernard et al., 1984

Residents of Yorkshire, England

 

 

Solvents (excluding benzene)

NA

0.45 (0.15–1.42)

 

Petrol products

NA

1.15 (0.37–3.56)

Hardell and Bengtsson, 1983

Male residents of three counties of Sweden

 

 

Low-grade exposure (age 25–85 years)

4

1.2 (0.4–3.8)

 

High-grade exposure (age 25–85 years)

14

3.0 (1.4–6.1)

 

Ever exposed to solvents or phenoxy acids or chlorophenols

7

6.6 (2.4–18.5)

NOTE: NA=not available.

a95% CIs were calculated by the committee with standard methods from the observed numbers presented in the original study.

bEstimated risk and 95% CIs were calculated by the committee with standard methods from the observed and expected numbers presented in the original study.

c90% CI is reported.

MULTIPLE MYELOMA

Description of Case-Control Studies

Table 6.37 describes the two case-control studies that the committee reviewed to assess the association between MM and exposure to organic solvents. Morris and colleagues (1986) evaluated the association between risk of MM and self-reported exposure to 20 chemicals that were included in a standard questionnaire administered to subjects or to family members if a subject was deceased or too ill to be interviewed. Using the same study population, Demers and co-workers (1993) evaluated the association between MM and employment in a variety of occupations and industries. Eriksson and Karlsson (1992) conducted a population-based case-control study in Sweden in which subjects or the next of kin of deceased subjects were interviewed with self-administered questionnaires and by telephone. Occupational and leisure-time exposure to organic solvents was reported by the subjects.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.37 Description of Case-Control Studies of Multiple Myeloma and Exposure to Organic Solvents

Reference

Description of Study Population

Number of Cases

Number of Controls

Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Morris et al., 1986

Cases identified through SEER tumor registries in four geographic areas who were under the age of 80 years and newly diagnosed in 1977–1981; population controls identified through random-digit dialing and a population survey of the four geographic areas

698

1,683

Aldehydes or ketones

Aromatic hydrocarbons

Chlorinated hydrocarbons

Paints or solvents

Interviews with a standardized questionnaire (direct or proxy) to assess occupational history (job titles) and specific exposure agents (self-reports); data obtained directly from 68% of cases and 99% of controls

Mantel-Haenszel

Sex, age, race, study area

 

Response rates: 89% of cases, 83% of controls

 

Eriksson and Karlsson, 1992

Cases reported to the Swedish Cancer Registry from the four northernmost counties of Sweden in 1982–1986; medical records were reviewed to confirm the diagnosis; for each living case, two controls were identified from the Swedish National Population Registry and cancer controls, and for each deceased case, one control was selected from the National Registry for Causes of Death

275

275

Organic solvents (occupational and leisure exposure)

Occupational histories (job titles) and specific leisure or work exposures obtained through questionnaire and supplemented by telephone interview (direct or proxy) (self-reports)

Univariate analysis

None

 

Response rates: 98.5% of cases, 95.3% of controls

 

Demers et al., 1993

Same study population as Morris et al., 1986 above.

692

1,683

Painters

Printing-machine operators

Same as above; occupations and industries coded according to the 1970 US Census

Mantel-Haenszel

Sex, age, race, study area

NOTE: SEER=Surveillance, Epidemiology, and End Results.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Epidemiologic Studies of Exposure to Organic Solvents and Multiple Myeloma

The association between MM and exposure to trichloroethylene was evaluated in cohort studies of workers who were monitored biologically for a metabolite of trichloroethylene (SIR=0.9, 95% CI=0.01–4.7) (Hansen et al., 2001), in trichloroethylene-production workers (SIR=0.57, 95% CI=0.01–3.17) (Axelson et al., 1994), in aircraft-manufacturing workers (SMR=0.91, 95% CI=0.34–1.99) (Boice et al., 1999), and in aircraft-maintenance workers (SMR=1.3, 95% CI=0.5–3.4) (Blair et al., 1998). In the Blair and colleagues’ study (1998), no increases in relative risk with increasing cumulative exposure were found, but women exposed intermittently to trichloroethylene at low levels had an RR of 4.4 (95% CI=1.0–20.4).

The authors of four cohort studies examined the relationship between exposure to benzene and the risk of MM. Rinsky and colleagues (1987) reported an SMR of 4.09 (95% CI=1.10–10.47) in the cohort of Pliofilm workers in Ohio. Additional years of followup did not add any cases of MM; with the increased number of person-years of observation, the strength of the association was reduced by about 70% (SMR=2.91, 95% CI=0.79–7.45) (Wong, 1995).

Among Monsanto chemical-plant workers who used benzene in various production operations, an RR of 3.23 (95% CI=0.7–9.4) was observed (Ireland et al., 1997), but no exposure-response pattern was found. A nested case-control study based on a cohort of Canadian petroleum-distribution workers showed an increased risk of MM only when benzene exposures were more than 20 ppm per year (OR=1.22, 95% CI=0.07–20.0; two cases) (Schnatter et al., 1996a), but there was no evidence of an exposure trend. In the Chinese study of benzene-exposed workers, no increased risk of MM was found (one exposed case; Yin et al., 1996a,b).

Cohort studies of aircraft workers and aircraft-maintenance workers assessed the association between risk of MM and exposure to solvents, other than trichloroethylene. Boice and colleagues (1999) found no evidence of an association with tetrachloroethylene. Blair and colleagues (1998) reported an increased risk of MM in those ever exposed to toluene (RRwomen=5.0, 95% CI=1.1–23.1), methyl ethyl ketone (RRwomen=4.6, 95% CI=0.9–23.2), or methylene chloride (RRmen=3.4, 95% CI=0.9–13.2).

A number of cohort and case-control studies were used to examine the association between exposures to unspecified mixtures of organic solvents and risk of MM, including several studies of painters (Boice et al., 1999; Demers et al., 1993; Lundberg and Milatou-Smith, 1998; Steenland and Palu, 1999), aircraft maintenance and manufacturing workers (Blair et al., 1998; Boice et al., 1999), and shoe manufacturers (Fu et al., 1996), all of whom used various solvent mixtures in their occupations. In one study (of shoe-manufacturing workers), benzene was a likely component of the mixtures, but relative risks were not specifically reported (Fu et al., 1996).

In a cohort study of paint manufacturers (Lundberg and Milatou-Smith, 1998), increased mortality from and incidence of MM were observed (SMR=3.8, 95% CI=1.0–9.7; SIR=3.2, 95% CI=0.9–8.3). Painters were at increased risk (SMR=1.70, 95% CI=0.46–4.35) in one study (Boice et al., 1999), but not in a study of painter-union members (SMR=0.97, 95% CI=0.75–1.24); it was similar to the risk found for nonpainters (SMR=0.90, 95% CI=0.52–1.46) (Steenland and Palu, 1999).

Other studies of MM and exposure to unspecified mixtures of organic solvents were conducted. Blair and colleagues (1998) found a 30% excess MM mortality among men (RR=1.3, 95% CI=0.4–3.8) and women (RR=1.9, 95% CI=0.4–8.2) exposed to any solvent. An

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

increased incidence of MM was found among a cohort of Swedish workers occupationally exposed to solvents (SIR=2.0, 95% CI=0.4–5.7) (Berlin et al., 1995). The study by Boice and colleagues (1999) did not show an association (other than in painters) between exposure to mixtures of solvents and risk of MM (SMR=1.17, 95% CI=0.95–1.45).

One case-control study (Demers et al., 1993; Morris et al., 1986) showed an increased relative risk of MM among painters. Morris and colleagues (1986) found an 80% increase in incidence of MM among subjects exposed to paints or mixtures of solvents (OR=1.8, 95% CI=1.2–2.7). Using the same population, Demers and colleagues (1993) observed a positive exposure-response pattern among painters according to length of employment (less than 10 years: OR=1.4, 95% CI=0.6–2.8; 10 years or more duration: OR=4.1, 95% CI=1.8–10.4).

Summary and Conclusion

None of the studies yielded persuasive evidence that exposure to trichloroethylene increased the risk of MM. The lack of increased relative risks was observed in the studies on exposure to benzene. Most studies on other specific solvents, such as tetrachloroethylene, toluene, methyl ethyl ketone, or methylene chloride, did not find a positive association, or if they did, they were not supported by other corroborating studies.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to specific solvents under review and multiple myeloma.

For exposure to solvent mixtures, a number of studies on painters found increased risk of MM, including one study that found an increasing risk with increasing years of employment (Demers et al., 1993). Table 6.38 identifies the key studies reviewed for each exposure and the data points evaluated by the committee. Unless indicated in the tables, the study populations include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is limited/suggestive evidence of an association between chronic exposure to solvents (as observed in studies of painters) and multiple myeloma.

TABLE 6.38 Selected Epidemiologic Studies—Multiple Myeloma and Exposures to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Trichloroethylene

Cohort Study—Incidence

Hansen et al., 2001

Biologically monitored Danish workers

 

 

Males, ever exposed

1

0.9 (0.01–4.7)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Routine exposure

6

0.91 (0.34–1.99)

 

Routine or intermittent

 

 

<1 year of exposure

3

0.45 (0.13–1.54)

 

1–4 years of exposure

8

1.48 (0.64–3.41)

 

≥5 years of exposure

3

0.51 (0.15–1.76)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Blair et al., 1998

Aircraft-maintenance workers in Utah

 

 

Any exposure (males and females)

14

1.3 (0.5–3.4)

 

Males

 

 

Low level intermittent

4

0.5 (0.1–1.9)

 

Low level continuous

4

0.8 (0.2–3.2)

 

Frequent peaks

4

0.8 (0.2–3.3)

 

Females

 

 

Low level intermittent

4

4.4 (1.0–20.4)

 

Low level continuous

1

2.4 (0.2–24.3)

 

Frequent peaks

1

0.9 (0.1–8.9)

 

Cumulative exposure, males (mortality)

 

 

No exposure

10

1.7 (0.5–5.5)

 

<5 unit-years

4

1.0 (0.2–4.2)

 

5–25 unit-years

2

0.8 (0.1–4.4)

 

>25 unit-years

4

1.2 (0.3–4.7)

 

Cumulative exposure, females (mortality)

 

 

No exposure

11

1.0 (0.1–9.9)

 

<5 unit-years

2

3.2 (0.5–19.8)

 

5–25 unit-years

1

4.3 (0.4–43.4)

 

≥25 unit-years

1

1.3 (0.1–13.2)

Axelson et al., 1994

Male workers in Sweden, ever exposed

1

0.57 (0.01–3.17)

Benzene

Cohort Studies—Mortality

Ireland et al., 1997

Male US chemical-plant workers

 

 

Production workers

 

 

Nonexposed

1

0.5 (0.0–2.8)

 

Any exposure

3

3.2 (0.7–9.4)

 

<12ppm-month

0

0.0 (0.0–10.1)

 

12–72 ppm-month

2

6.8 (0.8–2.5)

 

>72 ppm-month

1

3.7 (0.1–20.1)

Schnatter et al., 1996a

Male Canadian petroleum-distribution workers

 

 

0.00–0.49 ppm-years

3

1.00

 

0.50–7.99 ppm-years

1

0.39 (0.01–5.16)

 

8.00–1 9.99 ppm-years

1

0.60 (0.01–7.83)

 

20–219.8 ppm-years

2

1.22 (0.07–20.0)

Yin et al., 1996a,b

Chinese factory workers, ever exposed

1

0.4 (0.0–10.7)

Wong, 1995

Male US Pliofilm workers in Ohio (through 1987)

4

2.91 (0.79–7.45)

 

<40 ppm-years

3

3.21 (0.66–9.39)

 

40–200 ppm-years

0

0 (0–12.29)

 

200–400 ppm-years

0

0 (0–36.89)

 

>400 ppm-years

1

25.17 (0.63–139.83)

Rinsky et al., 1987

Male US Pliofilm workers in Ohio (through 1981)

4

4.09 (1.10–10.47)

 

<40 ppm-years

3

4.58 (0.92–13.39)

 

40–200 ppm-years

0

0

 

200–400 ppm-years

0

0

 

>400 ppm-years

1

53.47 (0.70–297.53)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Other Organic Solvents

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Tetrachloroethylene, routine exposure

1

0.40 (0.01–2.25)

 

Routine or intermittent exposure

 

 

<1 year of exposure

1

0.46 (0.06–3.48)

 

1–4 years of exposure

4

1.13 (0.38–3.35)

 

≥5 years of exposure

1

0.24 (0.03–1.84)

Blair et al., 1998

Aircraft-maintenance workers in Utah

 

 

Stoddard solvent

 

 

Males

96

1.0 (0.3–3.2)

 

Females

3

1.6 (0.3–8.2)

 

Isopropyl alcohol (males)

4

1.5 (0.4–6.4)

 

Trichloroethane (females)

2

13.2 (2.2–80.4)

 

Acetone

 

 

Males

4

1.6 (0.4–6.7)

 

Females

2

3.8 (0.6–23.8)

 

Toluene

 

 

Males

2

0.9 (0.2–4.8)

 

Females

4

5.0 (1.1–23.1)

 

Methyl ethyl ketone

 

 

Males

1

0.4 (0.1–4.0)

 

Females

3

4.6 (0.9–23.2)

 

Methylene chloride (males)

5

3.4 (0.9–13.2)

Unspecified Mixtures of Organic Solvents

Cohort Studies—Incidence

Lundberg and Milatou-Smith, 1998

Male Swedish paint-industry workers, ever employed

 

 

Incidence

4

3.2 (0.9–8.3)

Berlin et al., 1995

Swedish workers occupationally exposed to solvents

3

2.0 (0.4–5.7)

Lynge et al., 1995

Printing-industry workers in Denmark

 

 

Bookbinders, ever employed

1

1.27 (0.02–7.04)

 

Typographer (printing establishment), ever employed

4

1.14 (0.31–2.93)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Factory workers, ever exposed

90

1.17 (0.95–1.45)a

 

Mixed solvents, routine exposure

15

0.98 (0.55–1.61)

 

Routine or intermittent exposure

 

 

<1 year of exposure

3

0.34 (0.10–1.14)

 

1–4 years of exposure

13

0.63 (0.32–1.25)

 

≥5 years of exposure

34

0.80 (0.46–1.38)

 

Painter, employed >1 year

4

1.70 (0.46–4.35)

Steenland and Palu, 1999

US painters-union members

 

 

Painters

64

0.97 (0.75–1.24)

 

20-year membership

54

1.01 (0.76–1.32)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Lundberg and Milatou-Smith, 1998

Male Swedish paint-industry workers, ever employed

 

Mortality

4

3.8 (1.0–9.7)

Blair et al., 1998

Aircraft-maintenance workers in Utah

 

 

Any solvent

 

 

Males

19

1.3 (0.4–3.8)

 

Females

5

1.9 (0.4–8.2)

Fu et al., 1996

Shoe-manufacturing workers

 

 

English cohort

7

0.99 (0.40–2.05)

 

Probable solvent exposure

3

1.15 (0.24–3.36)

 

High solvent exposure

1

5.26 (0.13–29.30)

 

Florence cohort

3

3.70 (0.76–10.80)

 

Probable solvent exposure

1

2.17 (0.05–12.10)

 

High solvent exposure

1

2.44 (0.06–13.60)

Case-Control Studies

Demers et al., 1993

Residents of four US states

 

 

Painters

31

2.1 (1.2–3.6)

 

Employed <10 years

15

1.4 (0.6–2.8)

 

Employed ≥10 years

16

4.1 (1.8–10.4)

 

Printing machine operators

4

0.6 (0.1–2.0)

Eriksson and Karlsson, 1992

Residents of northern Sweden

 

Organic solvents (occupational), ever used

21

0.50 (0.30–0.80)b

 

Duration of exposure (occupational)

 

 

≤5 years

NA

0.37 (0.09–1.48)b

 

6–20 years

NA

0.59 (0.28–1.25)b

 

≥21 years

NA

0.37 (0.26–0.74)b

 

Organic solvents (leisure use), ever used

43

1.22 (0.80–1.89)b

Morris et al., 1986

Residents of four US states

 

 

Ever exposed to:

 

 

Aldehydes or ketones

7

1.1 (0.4–3.6)

 

Aromatic hydrocarbons

16

0.8 (0.5–1.4)

 

Chlorinated hydrocarbons

70

1.0 (0.7–1.4)

 

Paints or solvents

39

1.8 (1.2–2.7)

NOTE: NA=not available.

a95% CI calculated by the committee with standard methods from the observed and expected numbers presented in the original study.

b90% CI reported.

ADULT LEUKEMIA

With the introduction of the eighth and ninth revisions of the ICD codes and an increased ability to identify subtypes of leukemia, epidemiologists could more easily identify specific hematopoietic cancers in their studies. However, in most of the studies reviewed, especially those conducted early, it was not possible to identify specific subtypes of adult leukemia. Furthermore, in the small cohort studies, there were often insufficient cases of any particular type of leukemia, so all subtypes were combined in the analyses to improve statistical power. In contrast with the literature on insecticide exposure, there were enough studies on the specific subtypes of leukemia to present findings according to the following five groups: adult leukemia

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

broadly, acute leukemia, chronic leukemia, lymphatic leukemia, and hairy cell leukemia. The committee acknowledges that the literature on adult leukemia could be divided by specific cell type but believes that there were too few studies to support valid conclusions on each specific type of leukemia. The committee therefore determined that there were enough high quality studies to support conclusions regarding the broad category of adult leukemia and acute leukemia only, which includes acute myeloid leukemia (AML) and acute lymphocytic leukemia (ALL). The decision to group leukemia types by onset (acute vs chronic) rather than by cell type was based on the available literature. The studies on acute leukemia, including AML and ALL, appeared to share more risk factors and findings than the more limited literature on chronic myeloid leukemia (CML) and chronic lymphocytic leukemia (CLL). Thus, the literature directed the committee to draw conclusions of association that encompassed both subtypes of acute leukemia.

Description of Case-Control Studies

The characteristics of the case-control studies considered by the committee in drawing its conclusions of association are described below in Table 6.39. All but one case-control study (Aschengrau et al., 1993) included interviews with study subjects concerning occupational history. Several questionnaires also included queries about specific chemical exposures, including benzene, trichloroethylene, and tetrachloroethylene (Albin et al., 2000; Bernard et al., 1984; Clavel et al., 1996; Nordström et al., 1998; Richardson et al., 1992; Staines and Cartwright, 1993). In many studies, experts reviewed questionnaire responses and attributed exposures unaware of case or control status (Albin et al., 2000; Ciccone et al., 1993; Clavel et al., 1996, 1998; Costantini et al., 2001; Lazarov et al., 2000; Malone et al., 1989; Richardson et al., 1992). Another study used a job—exposure matrix (Clavel et al., 1998) to determine exposures. However, in some studies, self-reported exposures were used as the metric (Bernard et al., 1984; Clavel et al., 1995; Flodin et al., 1981; Mele et al., 1994; Nordström et al., 1998; Staines and Cartwright, 1993). Unlike most occupational studies of leukemia, the study by Aschengrau and colleagues (1993) assessed exposure to tetrachloroethylene on the basis of estimated concentrations in public drinking water in five towns of Cape Cod, Massachusetts. Case-control studies of leukemia and exposure to organic solvents that had reasonably good assessments of exposure and enough exposed cases include those by Albin and colleagues (2000), Clavel and colleagues (1996), Lazarov and colleagues (2000), Malone and colleagues (1989), and Richardson and colleagues (1992).

Epidemiologic Studies of Exposure to Organic Solvents and Adult Leukemia

Several large cohort studies provided evidence for evaluating the association between exposure to benzene and adult leukemia. The principal cohort studies included several occupational populations with estimated levels of exposure, including the Pliofilm workers (Crump, 1994; Rinsky et al., 1981, 1987), Chinese factory workers (Hayes et al., 1997; Yin et al., 1996a,b), US chemical workers (Bond et al., 1986; Ireland et al., 1997; Wong, 1987a,b), US and British petroleum-distribution and oil-refinery workers (Rushton and Alderson, 1981; Wong et al., 1993), and, in nested case-control studies, petroleum-distribution workers in Canada (Schnatter et al., 1996a,b) and the UK (Rushton and Romaniuk, 1997).

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.39 Description of Case-Control Studies of Leukemia and Exposure to Organic Solvents

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Viadana and Bross, 1972

White cases from the Tri-State Leukemia Survey, covering New York (New York City excluded), Baltimore, and Minnesota in 1959–1962; controls randomly selected from the same geographic areas

1345 leukemia

1,237

Painting work

Printing work

In-person interviews (direct or proxy) to assess occupational history (job titles)

OR

Age

Flodin et al., 1981

Deceased cases, identified through the records of the Linkoping University Hospital (Sweden) and parish registers, in 1972–1978; controls selected from parish registers, matched for age and sex

42 AML

244

Solvents

Mailed questionnaire to next of kin to assess occupational exposures (self-reports)

Unadjusted OR

None

Bernard et al., 1984

Cases identified through registries and clinician reporting among residents of the Yorkshire Health Region, UK, with diagnosis in 1979–1981 and hematologic confirmation; controls selected from hospital inpatients, matched for age, sex, and geographic area

79 lymphoid leukemia

79

Benzene

Solvents

In-person interview to assess occupational history (job titles) and details of solvent and chemical contacts (self-reports)

Logistic regression

Sex, age

Malone et al., 1989

Cases, age less than 80 years, diagnosed in four US SEER study areas in 1977–1981; controls selected through RDD or area sampling, depending on the study area, and matched for sex, race, and age

427 CLL

1683

Dry-cleaning industry work

Aromatic hydrocarbons

Paints

Interview (in-person or telephone) with standardized questionnaire to assess lifetime occupational and leisure exposure history (self-reports)

Logistic regression

Age, sex, race, education level, area of residence

 

Participation rates: 82.5% of cases, 83% of controls

 

Richardson et al., 1992

Cases, age 30 years or over, identified through hospital hematology departments in Paris and Creteil, France, in 1984–1988; hospital controls selected from other departments, matched for sex, age, ethnicity, and usual residence

185 acute leukemia

513

Benzene

Solvents

Other hydrocarbon solvents

Halogenated solvents

Oxygenated solvents

Interview with standardized questionnaire to assess lifetime occupational and leisure exposure history (self-reports) with solvent- exposure probability and level determined by industrial hygienist review

Conditional logistic regression

Matching variables, prior history of radiotherapy or chemotherapy

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Aschengrau et al., 1993

Cases reported to the Massachusetts Cancer Registry, diagnosed in 1983–1986 among residents of five upper Cape Cod towns; living controls were selected from the records of the HCFA and through RDD; deceased controls identified by the state Department of Vital Statistics and Research files

34 leukemia

737

Tetrachloroethylene

Relative delivered dose estimated in model accounting for location and years of residence, water flow, pipe characteristics

Logistic regression

Sex, age at diagnosis, vital status, educational level, occupational exposure to solvents, specific cancer risk factors

 

Response rates: 79.5% of cases, 75.9% of HCFA controls, 73.9% of RDD controls, 78.8% of next of kin of deceased controls

 

Ciccone et al., 1993

Cases, age 15–74 years, treated in the Main Hospital of Torino, Italy, newly diagnosed in 1989–1990; hospital controls selected in the same interval, matched for sex, age, and area of residence; population controls randomly selected from residents of Torino, matched on above

50 AML

17 CML

246

Benzene

Interview with standardized questionnaire to assess lifetime occupational history (job titles) with exposure to solvents determined by industrial hygienist

Logistic regression

Age, area of residence and of birth, smoking

 

Response rates: 91% of cases, 99% of hospital controls, 82% of population controls

 

Staines and Cartwright, 1993

Cases from the Yorkshire and Trent (UK) Regional Health Authority areas, identified through Leukemia Research Fund data collection or hematologist participation in 1985–1990 with histologic confirmation; hospital controls matched for sex, age, and hospital

50 HCL

95

Benzene

Solvents

Organic chemicals

In-person interview with structured questionnaire to assess occupational history and exposure to chemicals (self-reports)

Conditional logistic regression

Matching variables

Mele et al., 1994

Cases, age 15 years or over, identified by hematology departments in three Italian cities in 1986–1990; outpatients without hematologic disorders selected as controls

252 AML

100 ALL

156 CML

1,161

Painting work

Shoemakers

In-hospital interview to assess lifetime behavioral and occupational exposure histories (job titles; self-reports)

Logistic regression

Age, sex, education, residence outside study town, other occupations

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Clavel et al., 1995

Cases diagnosed in 18 French hospitals in 1980–1990; hospital controls selected from admission records during same interval, matched on sex, date of birth, date of admission, and residence (see also Clavel et al., 1996, 1998)

291 HCL

229 men

62 women

541

425 men

116 women

Solvents

Mailed questionnaire to assess lifetime occupational history (job titles) with supplementary telephone interview on exposure to solvents (self-reports)

Conditional logistic regression

Matching variables

Clavel et al., 1996

Male cases diagnosed in 18 French hospitals in 1980–1990; hospital controls selected from admission records during same interval, matched on sex, date of birth, date of admission, and residence (see also Clavel et al., 1995, 1998)

226 HCL

425

Benzene

Painters

Mailed questionnaire to assess lifetime occupational history (job titles) with supplementary telephone interview on exposure to solvents (self-reports) and use of job-exposure matrix

Conditional logistic regression

Matching variables, smoking, farm work

Clavel et al., 1998

Male cases diagnosed in 18 French hospitals in 1980–1990; hospital controls selected from admission records during same interval, matched on sex, date of birth, date of admission, and residence (see also Clavel et al., 1995, 1996)

226 HCL

425

Organic solvents

Launderers and dry cleaners

Painters

Printers

Spray painters

Mailed questionnaire to assess lifetime occupational history (job titles) with supplementary telephone interview on exposure to solvents (self-reports) and use of job-exposure matrix

Conditional logistic regression

Matching variables, smoking, farm work

Nordström et al., 1998

Male cases reported to the Swedish Cancer Registry in 1987–1992; controls selected from the National Population Registry, matched for age and county

111 HCL

400

Trichloroethylene

Acetone

White spirit

Solvents

Paint

Mailed questionnaire (with supplemental telephone followup) to assess lifetime occupational history (job titles) and exposure to solvents (self-reports)

Logistic regression

Age

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Albin et al., 2000

Consecutively diagnosed cases, age at least 20 years, at the Department of Clinical Genetics in Lund, Sweden, in 1976–1993 with hematologic confirmation; population-based controls selected through Statistics Sweden, matched for sex, year of birth, and county

333 AML

351

Benzene

Chlorinated solvents

Organic solvents

Telephone interview (direct or proxy) obtained lifetime occupational and leisure exposure histories (self-reports); team of occupational hygienists performed exposure assessment

Logistic regression

Age, sex

 

Participation rates: 90% of cases, 72% of controls

 

Lazarov et al., 2000

Consecutive white cases, age 18 years or over, identified through hematology departments in Novi Sad, Yugoslavia in 1986–1990, and London, England, in 1988–1994; controls selected from hospital inpatients, matched for hospital, admission year, sex, and age

98 AML

196

Solvents

Painting work

Paints

Oils

Machinery

Mechanics and fitters

In-person interview with questionnaire to assess occupational and exposure histories (self-reports)

Unadjusted OR, conditional logistic regression

Medication history, smoking status

Costantini et al., 2001

Cases, age 20–74 years, identified through periodic hospital survey and diagnosed in 12 regions of Italy in 1991–1993 with hematologic confirmation; controls randomly selected from municipal demographic files and the National Health Services files, matched for age and sex

652 leukemia

1,779

Launderers, dry cleaners, pressers Painting work

In-person interview (direct or proxy) with standardized and job-specific questionnaires to assess lifetime occupational history (job titles) and exposure to solvents (self-reports); probability of exposure further determined by industrial hygienist

Mantel-Haenszel OR

Age, sex

 

Response rates: 88% of cases, 81% of controls

 

NOTE: ALL=acute lymphatic leukemia; AML=acute myeloid leukemia; CLL=chronic lymphatic leukemia; CML=chronic myeloid leukemia; HCL=hairy cell leukemia; SEER=Surveillance, Epidemiology, and End Results; RDD=random-digit dialing; HCFA=Health Care Financing Administration.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

In the Pliofilm workers study (Crump, 1994; Rinsky et al., 1981, 1987), subjects included workers exposed to benzene for at least 1 day from 1940 to 1950. In the last followup study, there were 14 deaths from leukemia, resulting in an SMR of 2.9 (95% CI=1.61–4.95). The relative risk of leukemia was found to increase with cumulative exposure, regardless of the methods used to estimate levels of exposure. In the study by Crump (1994), the RR was 1.2 (95% CI=0.26–3.64) for exposure at 0–45 ppm per year and 3.1 for exposure at 400–1,000 ppm per year (95% CI=0.37–11.11); in the Rinsky and colleagues’ study (1987), the RR was 1.09 (95% CI=0.12–3.94) for 0.001–40 ppm per year and 3.22 (95% CI=0.36–11.65) for 40–200 ppm per year.

Yin and colleagues (1996a) investigated the relationship between leukemia and exposure to benzene among Chinese workers and reported increased leukemia mortality (RR=2.3, 95% CI=1.1–5.0) and incidence (RR=2.6, 95% CI=1.3–5.7). Hayes and colleagues (1997) evaluated the exposure-response relationships and did not find clear dose-response relationships, but they did find that the relative risks generally increased by year of hire, average exposure (ppm), duration of employment (years), and cumulative exposure (ppm-years).

Evidence of an association between exposure to benzene and leukemia was also provided by several other cohort studies, including studies of US chemical workers and petroleum-distribution workers. The studies tended to be small and lacked the statistical power to detect a small to moderate association. Specifically, findings from the US chemical-worker studies (Bond et al., 1986; Ireland et al., 1997; Wong, 1987a,b) showed increased risks but no clear dose-response relationships. Associations were found in the UK study of oil-refinery workers (SMR=2.33, 95% CI=0.98–5.56) (Rushton and Alderson, 1981) and the study of petroleum-marketing and -distribution workers (SMR=1.35, 95% CI=0.14–12.8 for at least 45 ppm per year), but no clear dose-response pattern emerged (Rushton and Romaniuk, 1997). In the study of Canadian petroleum-distribution workers (Schnatter et al., 1996a), no associations were found.

Associations between exposure to trichloroethylene and leukemia were not found in several cohort studies that had relatively accurate information on occupational exposures. They included studies of biologically monitored workers (Hansen et al., 2001), workers at a uranium-processing plant (Ritz, 1999), and aircraft and aerospace maintenance and manufacturing workers (Blair et al., 1998; Boice et al., 1999; Morgan et al., 1998). The studies were generally of high quality with sufficient statistical power to detect relative risks of leukemia.

Investigations of most other specific organic solvents were restricted to single studies, and estimates of relative risk were usually based on small numbers of exposed cases. As a result, in most studies, no associations with leukemia were reported (Boice et al., 1999; Dosemeci et al., 1991). However, in one study of Swedish rotogravure-plant workers (Svensson et al., 1990) exposed to toluene, an increased risk of leukemia was observed (SMR=1.67, 95% CI=0.34–4.88). The risk was higher in those exposed for 5 years or more and with at least a 10-year latency (SMR=2.54, 95% CI=0.52–7.42).

Occupations in which large quantities of organic solvents were used have been studied to identify health risks and instigate interventions to reduce exposure. In several studies reviewed by the committee, increased risks of leukemia were observed among painters (Costantini et al., 2001: OR=1.7, 95% CI=0.8–3.8; Lundberg and Milatou-Smith, 1998: SIR=1.5, 95% CI=0.3–4.3; Viadana and Bross, 1972: RR=2.82), car mechanics (Hunting et al., 1995: SMR=9.26, 95% CI=1.12–33.43), shoe-manufacturing workers (Fu et al., 1996: SMR=2.80, 95% CI =0.76–7.16; Paci et al., 1989: SMR=4.95, 95% CI=1.82–10.79), employees at a naval nuclear shipyard (Stern et al., 1986: OR=2.32, 95% CI=0.85–6.29), workers exposed to solvents

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

occupationally (Berlin et al., 1995: SIR=2.1, 95% CI=0.8–4.6), and leather or tannery workers (Costantini et al., 1989: SMR=1.64, 95% CI=0.53–3.82). Studies in which associations were not found (Anttila et al., 1998; Boice et al., 1999; Garabrant et al., 1988; Matanoski et al., 1986; Steenland and Palu, 1999; Walker et al., 1993; Wolf et al., 1981; Wong et al., 1993) generally had small numbers of exposed cases. Two studies (Anttila et al., 1995; Fu et al., 1996) provided evidence of a dose-response relationship with increasing relative risks as exposure increased.

Summary and Conclusion

IARC and the US Environmental Protection Agency (EPA) have determined that benzene is carcinogenic in humans on the basis of both animal and human studies (ATSDR, 1997a; IARC, 1987; NTP, 2001). That determination was based primarily on the findings on leukemia defined broadly. In addition, epidemiologic studies of occupations exposed to mixtures of solvents, including benzene have shown increased risks of developing cancer. Among those at risk are rubber workers, mechanics, and some groups of chemical workers, printers and paper-industry workers, and shoe and leather workers (IARC, 1987, 1989).

Based on its review of the literature on exposure to benzene, the committee found that the combination of consistently positive findings in the cohort of workers with known exposures to benzene and evidence of a dose-response relationship fulfilled the criteria for a conclusion of sufficient evidence of an association between exposure to benzene and adult leukemia. However, the committee decided that the evidence of an association between exposure to benzene and adult leukemia was not as strong as that for acute leukemia. Thus, it did not warrant a conclusion of causality. The findings, although mostly positive, are not as consistent and statistically precise as the findings on acute leukemia. Most likely, the positive studies on adult leukemia and exposure to benzene include cases of acute leukemia. However, they may also include cases of chronic leukemia, lymphatic leukemia, and hairy cell leukemia, for which the existence of associations is not as clear. On the basis of the studies reviewed, the committee believes that the evidence on exposure to benzene and adult leukemia, defined broadly, met the definition of sufficient evidence of an association but not sufficient evidence of a causal relationship.

The committee concludes, from its assessment of the epidemiologic literature, that there is sufficient evidence of an association between chronic exposure to benzene and adult leukemia.

For exposure to other solvents, such as trichloroethylene and toluene, the overall paucity of studies and the lack of consistently positive findings limits the evidence that the committee had to review.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to specific organic solvents under review, other than benzene, and adult leukemia.

In contrast, the findings for unspecified mixtures of organic solvents and adult leukemia showed increased relative risks, including two studies that provided evidence for a dose-response relationship with increasing levels of exposure. Table 6.40 identifies the key studies reviewed by the committee on adult leukemia. Unless indicated in the table, the study populations include both men and women.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

The committee concludes, from its assessment of the epidemiologic literature, that there is limited/suggestive evidence of association between chronic exposure to unspecified mixtures of organic solvents and adult leukemia.

TABLE 6.40 Selected Epidemiologic Studies—Adult Leukemia and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Benzene

Cohort Studies—Mortality

Crump, 1994

Male US Pliofilm workers

14

2.9 (1.61–4.95)a

 

Cumulative exposure

 

 

0–45 ppm-years

3

1.2 (0.26–3.64)a

 

45–400 ppm-years

4

2.7 (0.73–6.83)a

 

400–1,000 ppm-years

2

3.1 (0.37–11.11)a

 

>1000 ppm-years

5

28.1 (9.00–64.83)a

Yin et al., 1996a

Chinese factory workers

 

 

Mortality, total

38

2.3 (1.1–5.0)

 

Males

25

2.1 (1.0–5.3)

 

Females

13

2.8 (0.8–17.6)

 

 

42

 

 

Incidence, total

 

2.6 (1.3–5.7)

Hayes et al., 1997

Chinese factory workers

 

 

All exposed workers

38

2.5 (1.2–5.1)

 

Exposed workers, year of hire

 

 

<1972

25

2.4

 

≥1972

13

3.4

 

Exposed workers, average ppm

 

 

<10

15

2.0 (0.9–4.5)

 

10–24

13

3.7 (1.6–8.7)

 

≥25

10

2.3 (0.9–5.7)

 

p trend=0.02

 

Exposed workers, duration

 

 

<5 years

14

4.0 (1.7–9.6)

 

5–9 years

11

3.1 (1.3–7.5)

 

≥10 years

13

1.5 (0.6–3.6)

 

p trend=0.98

 

Exposed workers, cumulative exposure

 

 

<40 ppm-years

11

1.9 (0.8–4.7)

 

40–99 ppm-years

8

3.1 (1.2–8.0)

 

≥100 ppm-years

19

2.7 (1.2–6.0)

 

p trend=0.04

Ireland et al., 1997

Male US chemical-plant workers

 

 

<12 ppm-months

2

2.5 (0.3–8.9)

 

12–72 ppm-months

0

0.0 (0.0–5.4)

 

≥72 ppm-months

3

4.6 (0.9–13.4)

Rushton and Romaniuk, 1997

Male UK petroleum marketing and distribution workers

 

Cumulative exposure

 

 

<0.45 ppm-years

22

1.00

 

0.45–4.49 ppm-years

47

1.42 (0.77–2.61)

 

4.5–44.9 ppm-years

20

2.48 (0.73–3.00)

 

≥45 ppm-years

1

1.35 (0.14–12.8)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Schnatter et al., 1996a

Male Canadian petroleum distribution workers

 

 

By median, 75th, and 90th percentiles

 

 

0.0–0.49 ppm-years

10

1.00

 

0.50–7.99 ppm-years

1

0.22 (0.0–1.82)

 

8.0–19.99 ppm-years

1

0.42 (0.01–3.95)

 

20.0–219.8 ppm-years

2

0.96 (0.09–6.81)

 

Exposure by regulatory standards

 

 

0.00–0.45 ppm-years

10

1.00b

 

>0.45–4.5 ppm-years

1

0.43 (0.0–4.05)

 

4.5–45 ppm-years

1

0.16 (0.0–4.55)

 

≥45 ppm-years

2

1.47 (0.16–13.1)

Rinsky et al., 1987

Male US Pliofilm workers

 

 

Ever exposed

9

3.37 (1.54–6.41)

 

0.001–40 ppm-years

2

1.09 (0.12–3.94)

 

40–200 ppm-years

2

3.22 (0.36–11.65)

 

200–400 ppm-years

2

11.86 (1.33–42.85)

 

>400 ppm-years

3

66.37 (13.34–193.9)

Wong, 1987a,b

Male chemical manufacturing workers

 

 

Entire cohort

7

0.75 (0.30–1.54)

 

Continuously exposed

6

1.35 (0.49–2.95)

 

Cumulative exposure

 

 

<180 ppm-months

2

0.97 (0.12–3.49)

 

180–719 ppm-months

1

0.78 (0.20–4.34)

 

≥720 ppm-months

3

2.76 (0.57–8.06)

Bond et al., 1986

Male chemical workers, ever exposed

 

 

Cohort less those exposed to arsenic, asbestos, or high levels of vinyl chloride

3

1.62 (0.33–4.61)

Rushton and Alderson, 1981

Male employees in eight UK oil refineries, ever exposed

NA

2.33 (0.98–5.56)

Trichloroethylene

Cohort Study—Incidence

Hansen et al., 2001

Biologically monitored Danish workers

 

 

Males

5

1.9 (0.6–4.4)

 

Females

1

3.1 (0.04–18.0)

Cohort Studies—Mortality

Ritz, 1999

White male employees at a uranium-processing plant in Ohio

12

1.09 (0.56–1.91)

Boice et al., 1999

Aircraft-manufacturing workers in California, ever exposed

2

1.05 (0.54–1.84)

Blair et al., 1998

Aircraft-maintenance workers in Utah

 

 

Combined early and recent followup cohort

16

0.6 (0.3–1.2)

 

Men (cumulative exposure)

 

 

No exposure

9

1.0 (0.4–2.9)

 

<5 unit-years

7

1.0 (0.3–3.2)

 

5–25 unit-years

0

 

>25 unit-years

7

1.2 (0.4–3.6)

Morgan et al., 1998

Aerospace workers in Arizona, ever exposed

10

1.05 (0.50–1.93)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Other Specific Organic Solvents

Cohort Study—Incidence

Svensson et al., 1990

Male rotogravure workers in Sweden—toluene

 

 

All exposed

3

1.67 (0.34–4.88)

 

≥5 years of exposure and >10-year latency

3

2.54 (0.52–7.42)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Ever exposed—tetrachloroethylene

5

1.09 (0.35–2.55)

Dosemeci et al., 1991

Male workers exposed to phenol

 

 

No exposure

7

0.9

 

All exposed

14

0.9 (0.5–1.4)

 

High exposure

1

0.8

 

Medium exposure

5

0.7

 

Low exposure

8

1.0

Case-Control Studies

Costantini et al., 2001

Residents of 12 areas of Italy

 

 

Launderers, dry cleaners, pressers, ever employed

2

3.3 (0.3–32.4)

Aschengrau et al., 1993

Residents of Cape Code, MA—tetrachloroethylene

 

 

Exposure history

 

 

Any

7

1.77 (0.63–4.33)

 

Low

5

1.38 (0.40–3.78)

 

High

2

5.95 (0.58–31.72)

Unspecified Mixtures of Organic Solvents

Cohort Studies—Incidence

Anttila et al., 1998

Finnish workers monitored for exposure to solvents

 

 

Aromatic hydrocarbons, all years

1

0.30 (0.01–1.67)

Lundberg and Milatou-Smith, 1998

Swedish paint-industry workers

 

Incidence, male workers, ≥5 years

3

1.5 (0.3–4.3)

Berlin et al., 1995

Swedish workers occupationally exposed to solvents

 

 

Incidence

6

2.1 (0.8–4.6)

Anttila et al., 1995

Finnish workers exposed to halogenated hydrocarbons

 

 

Ever exposed, male and female for the entire measurement period

5

1.08 (0.35–2.53)

 

Mean personal level

 

 

<100 µmol/L

1

0.39 (0.01–2.19)

 

100+ µmol/L

4

2.65 (0.72–6.78)

Cohort Studies—Mortality

Boice et al., 1999

Aircraft-manufacturing workers in California

 

 

Mixed solvents, ever exposed

28

1.02 (0.68–1.48)

Steenland and Palu, 1999

Painter-union members

 

 

Painters (all members)

138

0.92 (0.78–1.11)

 

Painters (20 years since first union membership)

111

1.11 (0.74–1.08)

Lundberg and Milatou-Smith, 1998

Swedish paint-industry workers

 

Mortality, male workers, ≥5 years

2

1.2 (0.2–4.3)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Fu et al., 1996

Shoe-manufacturing workers

 

 

English cohort

16

0.89 (0.51–1.45)

 

Probable solvent

4

0.68 (0.19–1.75)

 

High solvent

0

 

Florence cohort

8

2.14 (0.92–4.21)

 

Probable solvent

4

2.52 (0.69–6.44)

 

High solvent

4

2.80 (0.76–7.16)

Hunting et al., 1995

Male vehicle mechanics in the District of Columbia

 

 

Solvents or fuels, high exposure

2

9.26 (1.12–33.43)

Berlin et al., 1995

Swedish workers occupationally exposed to solvents

 

 

Mortality

4

2.3 (0.6–5.8)

Walker et al., 1993

Shoe-manufacturing workers, employed >1 month

15

1.11 (0.63–1.85)

Wong et al., 1993

Gasoline distribution workers in the United States

 

 

Land-based terminal cohort

27

0.89 (0.59–1.29)

 

Marine-based terminal cohort

16

0.70 (0.42–1.09)

Costantini et al., 1989

Male tannery workers in Tuscany, employed >6 months

5

1.64 (0.53–3.82)

Paci et al., 1989

Shoe factory workers in Italy

 

 

Total years of exposure (males)

6

4.95 (1.82–10.79)a

Garabrant et al., 1988

Aircraft manufacturing workers in California

 

 

Employed >4 years

16

0.82 (0.47–1.34)

Matanoski et al., 1986

Painters and allied tradesmen union members

44

0.93 (0.68–1.25)b

Stern et al., 1986

Male employees at a naval nuclear shipyard

 

 

Ever worked in job with solvent exposure

NA

2.32 (0.85–6.29)

 

8.59 years in a solvent job

NA

1.82 (0.93–3.58)

Wolf et al., 1981

Workers in the US rubber industry

 

 

High solvent

8

0.8 (p=0.64)

 

Medium solvent

38

1.1 (p=0.79)

 

Low solvent

23

0.6 (p=0.05)

Alderson and Rattan, 1980

Male chemical-plant workers

 

Methyl ethyl ketone plant production workers

1

2.86 (0.07–15.91)a

Chiazze et al., 1980

Male spray painters at automobile-assembly plants

2

1.13 (0.14–4.08)b

Case-Control Studies

Costantini et al., 2001

Residents of 12 areas of Italy

 

 

Painters, ever employed

10

1.7 (0.8–3.8)

Viadana and Bross, 1972

Residents of New York, Baltimore, and Minnesota

 

 

(Tri-State Leukemia Survey)

 

 

Painters, white males

 

 

15–44 years

4

2.20

 

45–64 years

12

3.29

 

65–48 years

15

2.90

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

 

Printers, white males

 

 

15–44 years

2

8.50

 

45–64 years

4

1.40

 

65–48 years

11

1.00

NOTE: NA=not available.

a95% CI were calculated by the committee with standard methods from the observed and expected numbers presented in the original papers.

bEstimated risks and 95% CI were calculated by the committee with standard methods from the observed and expected numbers presented in the original papers.

Epidemiologic Studies of Exposure to Organic Solvents and Acute Leukemia

Virtually all studies of exposure to benzene and acute leukemia (specifically AML or acute nonlymphocytic leukemia [ANLL], and ALL) showed positive associations (Albin et al., 2000; Ciccone et al., 1993; Crump, 1994; Ireland et al., 1997; Richardson et al., 1992; Rushton and Romaniuk, 1996; Wong, 1995; Yin et al., 1996a,b). In several cohort and case-control studies, the risks of acute leukemia were found to increase as the level or duration of exposure to benzene increased (Ireland et al., 1997; Richardson et al., 1992; Rushton and Romaniuk, 1996; Wong, 1995).

Increases in incidence of and mortality from AML were reported in the cohort studies of Chinese factory workers (SMR=3.1, 95% CI=1.2–10.7) (Yin et al., 1996a,b) and Pliofilm workers in Ohio (SMR=5.03, 95% CI=1.84–10.97) (Wong, 1995) respectively. Increased risks of AML were also reported in the cohort of petroleum-distribution workers in the UK (Rushton and Romaniuk (1996) and in a case-control study in Sweden by Albin and colleagues (2000) (OR=1.5, 95% CI=0.89–2.6 for all exposure levels). In a case-control study in France, Richardson and colleagues (1992) found an increased odds ratio among those with “medium/high” exposure to benzene (OR=2.8, 95% CI=1.3–5.9) when other occupational exposures were adjusted for. Increased risk of ANLL (RR=1.4, 95% CI=0.4–3.42) was reported from the cohort study of Monsanto chemical workers (Ireland et al., 1997). Yin and colleagues (1996a,b) found increased ALL risk among benzene-exposed workers, although the CIs were quite broad (SIR=2.8, 95% CI=0.5–54.5). Weak positive associations were observed in a low-powered case-control study in which exposure to solvents was assessed by industrial hygienists (Ciccone et al., 1993).

A number of case-control studies showed mostly positive associations between acute leukemia and exposure to unspecified mixtures of organic solvents (Albin et al., 2000; Flodin et al., 1981; Lazarov et al., 2000; Mele et al., 1994; Richardson et al., 1992). To increase the accuracy of exposure and to reduce recall bias, industrial hygienists, unaware of the case status of subjects, attributed exposure on the basis of job descriptions (Albin et al., 2000; Lazarov et al., 2000; and Richardson et al., 1992). In a case-control study by Albin and colleagues (2000), the OR for AML for all levels of exposure to organic solvents was 1.6 (95% CI=1.1–2.4) and 2.3 (95% CI=1.0–5.0) for moderate to high levels of exposure. For exposure to solvents in general, Lazarov and colleagues (2000) also showed an increased risk of AML (OR=2.52, 95% CI=1.45–4.39); the risk increased to 3.86 (95% CI=1.83–8.14) for “probable exposure” to solvents. In the case-control study by Flodin and colleagues (1981), the risk of AML increased to 6.3 (95% CI=2.6–15.3). A study of people ever employed as painters or shoemakers showed increased relative risks of both AML and ALL: for painters, the OR was 3.2 (95% CI=0.5–20.8)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

for AML and 4.7 (95% CI=0.6–34.2) for ALL; for shoemakers, the OR was 2.4 (95% CI=0.9–6.9) for AML and 1.3 (95% CI=0.2–10.2) for ALL.

Summary and Conclusion

On the basis of the consistently high relative risks in studies in which the exposure to benzene is well known, the committee decided that the evidence meets the requirement for a conclusion of causality between chronic exposure to benzene and acute leukemia. Furthermore, given the strong positive associations in the cohort studies of highly exposed subjects, it is likely that confounding and selection bias do not account for the findings. Experimental evidence supports a biologic mechanism that strengthens the conclusion. The details of that experimental evidence are provided in Chapter 4 and discussed below.

The metabolism of benzene, which occurs in the liver and to a smaller extent in the bone marrow, plays an important role in its toxicity. Benzene is metabolized to benzene oxide, an epoxide, through an oxidation reaction catalyzed primarily by cytochrome P450 2E1. Benzene oxide can be metabolized to various compounds, including o-benzoquinone and p-benzoquinone, which are thought to be the two main metabolites that mediate the toxicity of benzene. Data on laboratory animals show that benzene affects the bone marrow in a dose-dependent manner, causing anemia, leukopenia, and thrombocytopenia; continued exposure causes aplasia and pancytopenia (Bruckner and Warren, 2001). Benzene also has carcinogenic properties. In experimental animals, increases in incidence of malignant lymphoma and some solid tumors have been seen after exposure to high doses of benzene.

The committee concludes, from its assessment of the epidemiologic and experimental literature, that there is sufficient evidence of a causal relationship between chronic exposure to benzene and acute leukemia.

For exposure to unspecified mixtures of organic solvents, the studies were virtually all positive, including several that were statistical strong. One study provided evidence of an exposure-response relationship in terms of both increasing levels and increasing duration of exposure. Table 6.41 identifies all the studies reviewed by the committee. Unless indicated in the table, the study populations include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is sufficient evidence of an association between chronic exposure to unspecified mixtures of organic solvents and acute leukemia.

TABLE 6.41 Selected Epidemiologic Studies—Acute Leukemia and Exposure to Organic Solvents

Reference

Study Population and Cancer Type

Exposed Cases

Estimated Relative Risk (95% CI)

Benzene

Cohort Studies—Mortality

Ireland et al., 1997

Male Monsanto Company production workers in Sauget, Illinois-ANLL

 

 

<12ppm-months

1

3.7 (0.1–20.6)

 

12–72 ppm-months

0

0.0 (0.0–44.1)

 

>72 ppm-months

1

4.5 (0.1–25.3)

 

Ever exposed

4

1.4 (0.4–3.42)a

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population and Cancer Type

Exposed Cases

Estimated Relative Risk (95% CI)

Rushton and Romaniuk, 1996

Male petroleum-distribution workers in the UK—AML (myeloid and monocytic)

 

 

<0.45 ppm-years

7

1

 

0.45–4.49 ppm-years

15

2.17 (0.77–6.09)

 

4.5–44.9 ppm-years

9

2.82 (0.82–9.38)

 

>45 ppm-years

Yin et al., 1996a,b

Chinese factory workers, ever exposed

 

 

Incidence of AML

23

3.1 (1.2–10.7)

 

Incidence of ALL

5

2.8 (0.5–54.5)

Wong, 1995

Male Pliofilm workers in Ohio—AML (myeloid and monocytic)

 

 

<40 ppm-years

1

1.19 (0.03–6.63)

 

40–200 ppm-years

0

0 (0–14.75)

 

200–400 ppm-years

2

27.21 (3.29–98.24)

 

>400 ppm-years

3

98.37 (20.28–287.65)

 

Total

6

5.03 (1.84–10.97)

Crump, 1994

Male Pliofilm workers in Ohio—AML (myeloid and monocytic)

 

 

Total (two cases could not be identified as to type)

8–10

5.0–6.2

 

0–45 ppm-years)

0–2

0.0–2.4

 

45–400 ppm-years

1

2.0 (0.05–10.92)b

 

400–1000 ppm-years

2

9.1 (1.10–32.82)b

 

>1000 ppm-years

5

82.8 (27.00–194.50)b

Case-Control Studies

Albin et al., 2000

Residents of Lund, Sweden—AML

 

 

All exposure levels

39

1.5 (0.89–2.6)

 

Hobby—low levels

28

1.6 (0.89–3.1)

 

Moderate-high levels

11

1.3 (0.52–3.1)

Ciccone et al., 1993

Residents in the Main Hospital of Torino, Italy—AML, CML, MDS; ever exposed

10

1.7 (0.6–5.5)

Richardson et al., 1992

Residents of Paris and Créteil, France—acute leukemia; occupationally exposed

 

 

All exposure levels

22

1.3 (0.8–2.3)

 

High or medium exposure

15

2.8 (1.3–5.9)

Unspecified Mixtures of Organic Solvents

Case-Control Studies

Albin et al., 2000

Residents of Lund, Sweden—AML

 

 

Organic solvents, all exposure levels

88

1.6 (1.1–2.4)

 

Low levels (hobby use)

70

1.5 (1.0–2.3)

 

1–7 years-duration

10

0.72 (0.32–1.6)

 

8–14 years-duration

18

1.5 (0.73–3.1)

 

15–20 years-duration

42

2.1 (1.2–3.7)

 

Moderate-high levels

18

2.3 (1.0–5.0)

 

Moderate for 1–20 years

9

1.6 (0.57–4.3)

 

High for 1–20 years

9

3.9 (1.0–15)

 

Chlorinated solvents

12

0.78 (0.36–1.7)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population and Cancer Type

Exposed Cases

Estimated Relative Risk (95% CI)

Lazarov et al., 2000

Residents of Novi Sad, Yugoslavia, and London, England—AML

 

 

Solvents (multivariate analysis)

53

2.52 (1.45–4.39)

 

Possible exposure

24

2.28 (1.12–4.62)

 

Probable exposures

29

3.86 (1.83–8.14)

 

Painters and related workers

7

4.57 (1.29–16.14)

 

Paints

23

1.28 (0.68–2.43)

 

Oils

32

1.56 (0.87–2.81)

 

Machinery mechanics and fitters

12

4.03 (1.44–11.23)

Mele et al., 1994

Residents of Italy

 

 

Painters, ever employed

 

 

AML

4

3.2 (0.5–20.8)

 

ALL

4

4.7 (0.6–34.2)

 

Shoemakers, ever employed

4

2.4 (0.9–6.9)

 

AML

1

1.3 (0.2–10.2)

 

ALL

 

Richardson et al., 1992

Residents of Paris and Créteil, France—acute leukemia; ever exposed

 

 

Solvents, all exposure levels

71

1.1 (0.7–1.5)

 

Other hydrocarbon solvents, all exposure levels

28

1.0 (0.6–1.6)

 

High or medium exposure

11

0.9 (0.4–1.8)

 

Halogenated solvents, all exposure levels

44

1.1 (0.7–1.6)

 

High or medium exposure

23

1.0 (0.6–1.7)

 

Oxygenated solvents, all exposure levels

42

1.5 (1.0–2.4)

 

High or medium exposure

15

1.5 (0.7–2.7)

Flodin et al., 1981

Residents of Linköping, Sweden—AML

 

 

Solvents, ever exposed

11

6.3 (2.6–15.3)b

aRR and 95% CI calculated by the committee with standard methods from the observed and expected numbers presented in the original study.

b95% CI calculated by the committee with standard methods from the observed and expected numbers presented in the original study.

Epidemiologic Studies of Exposure to Organic Solvents and Chronic Leukemia

An association between exposure to benzene and risk of chronic leukemia was reported in several studies (Ireland et al., 1997: SMR=2.3, 95% CI=0.7–5.3; Rushton and Romaniuk, 1997: SMR=1.22, 95% CI=0.38–3.89 for 4.5–44.9 ppm-years; Yin et al., 1996a,b: SMR=2.6, 95% CI=0.7–16.9). No excess risks (OR=1.1, 95% CI=0.6–2.0) were observed for self-reported exposure to aromatic hydrocarbons (the class of solvents that includes benzene, toluene, and xylene) in a population-based case-control study of CLL (Malone et al., 1989). Except for one exposure level, the case-control study by Rushton and Romaniuk (1997) showed relative risks close to unity.

A case-control study of leukemia and preleukemia conducted in Italy found that shoemakers and painters were at high risk for CML (OR=4.5, 95% CI=1.6–13.0 for shoemakers; OR=7.6, 95% CI=1.5–39.8 for painters) (Mele et al., 1994). Positive associations between exposure to mixed solvents and CLL were found in a US study reporting previous exposures to paints (OR=1.4, 95% CI=0.8–2.2) and hobby painting (OR=1.4, 95% CI=0.9-

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

2.0), but no increased risks were found (OR=0.9) among those employed in the dry-cleaning industry (Malone et al., 1989).

The committee drew no conclusion on an association between exposure to benzene or unspecified mixtures of organic solvents and chronic leukemia. Table 6.42 identifies the studies reviewed by the committee on chronic leukemia. Unless indicated in the table, the study populations include both men and women.

TABLE 6.42 Selected Epidemiologic Studies—Chronic Leukemia and Exposure to Organic Solvents

Reference

Study Population and Cancer Type

Exposed Cases

Estimated Relative Risk (95 % CI)

Benzene

Cohort Studies—Mortality

Ireland et al., 1997

Male Monsanto plant workers in Sauget, IL

 

 

Production workers exposed to benzene (all leukemias)

5

2.3 (0.7–5.3)

 

CLL (exposure to benzene category)

 

 

<12 ppm-months

1

5.9 (0.1–32.6)

 

12–72 ppm-months

0

0.0 (0.0–24.7)

 

>72 ppm-months

1

6.7 (0.2–37.7)

Rushton and Romaniuk, 1997

Male UK petroleum marketing and distribution workers CLL (exposure to benzene)

 

<0.45 ppm-years

8

1

 

0.45–4.49 ppm-years

16

1.07 (0.40–2.86)

 

4.5–44.9 ppm-years

7

1.22 (0.38–3.89)

 

≥45 ppm-years

0

0

 

CML (exposure to benzene)

 

 

<0.45–4.49 ppm-years

9

1

 

4.5–44.9 ppm-years

2

0.76 (0.14–4.06)

 

≥45 ppm-years

0

Yin et al., 1996a,b

Chinese factory workers—CML

 

 

Any exposure

9

2.6 (0.7–16.9)

Case-Control Study

Malone et al., 1989

US residents of four areas—CLL

 

 

Aromatic hydrocarbons, any exposure

26

1.1 (0.6–2.0)

Unspecified Mixtures of Organic Solvents

Case-Control Studies

Mele et al., 1994

Residents of Italy

 

 

Painters, ever employed—CML

5

7.6 (1.5–39.8)

 

Shoemakers, ever employed—CML

6

4.5 (1.6–13.0)

Malone et al., 1989

US residents of four areas—CLL

 

 

Paints, ever exposed

26

1.4 (0.8–2.2)

 

Dry-cleaning industry, ever employed

14

0.9 (0.4–1.8)

 

Hobby painting, ever exposed

48

1.4 (0.9–2.0)

Epidemiologic Studies on Exposure to Organic Solvents and Lymphatic Leukemia

Several studies of lymphatic leukemia examined associations with exposure to benzene. Positive associations between lymphatic leukemia and exposure to benzene were observed in several studies (Bernard et al., 1984; Wilcosky et al., 1984), but there was considerable variability in the estimates.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Wilcosky found increased risk estimates associated with almost all exposures, including ethyl acetate (OR=5.3, p<0.05) and acetone (OR=6.8, p<0.01); this study was a large-scale investigation of rubber workers in which individual estimates for a variety of solvents were made on the basis of company records. Bernard and colleagues (1984) found a positive association between lymphatic leukemia and occupational solvents, excluding benzene (OR=1.56, 95% CI =0.47–5.17). Among rubber workers, relative risks increased with increasing estimated levels of exposure to mixed solvents (McMichael et al., 1975; Wolf et al., 1981). A cohort study of lymphatic leukemia in nuclear-plant workers found increased risk associated with jobs entailing exposure to unspecified mixtures of organic solvents (RR=1.99, 95% CI=0.46–8.67) (Stern et al., 1986).

No conclusions were drawn on an association between exposure to benzene, other specific organic solvents, or unspecified mixtures of solvents and lymphatic leukemia. Table 6.43 identifies the key studies on lymphatic leukemia reviewed by the committee. Unless indicated in the table, the study populations include both men and women.

TABLE 6.43 Selected Epidemiologic Studies—Lymphatic Leukemia and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Benzene

Cohort Studies—Mortality

Wilcosky et al., 1984

White, male US rubber workers, exposed >1 year

4

2.5

Arp et al., 1983

US rubber workers

 

 

Primary exposure to benzene

3

4.50

 

Secondary exposure to benzene

2

1.50

Case-Control Study

Bernard et al., 1984

Male residents of Yorkshire, England, ever exposed

NA

3.06 (0.98–11.97)

Other Specific Solvents

Cohort Study—Mortality

Wilcosky et al., 1984

White, male US rubber workers, exposed >1 year

 

 

Acetone

3

6.8

 

Ethanol

4

2.0

 

Ethyl acetate

3

5.3

 

Isopropanol

6

1.8

 

Solvent “A” (mixture of toluene and other solvents)

7

2.8

 

Specialty naphthas

8

2.8

 

Perchloroethylene (tetrachloroethylene)

1

 

Phenol

1

 

Toluene

2

3.0

 

Trichloroethylene

2

0.81

 

VM and P naphthas

3

2.9

 

Xylenes

4

3.3

Unspecified Mixtures of Solvents

Cohort Studies—Mortality

Stern et al., 1986

Male Portsmouth Naval Shipyard workers

 

 

Solvent job

NA

1.99 (0.46–8.67)

Arp et al., 1983

US rubber workers

 

 

Primary exposure to other solvents

11

4.50

 

Secondary exposure to other solvent

8

1.60

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Wolf et al., 1981

US rubber workers

 

 

Low exposure

11

0.8

 

Medium exposure

16

1.2

 

High exposure

5

1.6

McMichael et al., 1975

Male US rubber workers

 

 

All exposure levels

NA

3.25

 

Low exposure

NA

2.50

 

Medium exposure

NA

2.00

 

High exposure

NA

5.50

 

p-trend<0.025

Case-Control Study

Bernard et al., 1984

Male residents of Yorkshire, England

 

 

Solvents (excluding benzene use)

NA

1.56 (0.47–5.17)

NOTE: NA=not available.

Epidemiologic Studies of Exposure to Organic Solvents and Hairy Cell Leukemia

In a study by Clavel and colleagues (1995, 1996, 1998), no excess risks were observed between exposure to benzene and hairy cell leukemia. In another case-control study, exposure to benzene was associated with risk of hairy cell leukemia (Staines and Cartwright, 1993), although CIs included unity (OR=2.00, 95% CI=0.50–8.00).

Nordström and colleagues (1998) found an association between hairy cell leukemia and self-reported exposures to white spirit (or naphtha) (OR=2.0, 95% CI=1.2–3.4), acetone (OR=1.2, 95% CI=0.3–4.3), and trichloroethylene (OR=1.5, 95% CI=0.7–3.3).

Clavel and colleagues (1995, 1998) inquired about exposure to organic solvents and found little evidence of an association between hairy cell leukemia and exposure to solvent mixtures or occupations involving exposure to organic solvents, such as painting, spray painting, printing, and laundry and dry cleaning. Staines and Cartwright (1993) also did not find strong evidence of an association with organic solvents (OR=1.45, 95% CI=0.58–3.66), while Nordström and colleagues (1998) found positive associations between hairy cell leukemia and exposure to all solvents (OR=1.5, 95% CI=0.99–2.3) and working with paints (OR=4.3, 95% CI=1.8–10.3).

No conclusions were drawn on an association between exposure to specific solvents or unspecified mixtures of organic solvents and hairy cell leukemia. Table 6.44 identifies the studies related to hairy cell leukemia and solvent exposure. Unless indicated in the table, the study populations include both men and women.

TABLE 6.44 Selected Epidemiologic Studies—Hairy Cell Leukemia and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Benzene

Case-Control Studies

Clavel et al., 1996

Male residents of France

 

 

All exposures to benzene

34

0.8 (0.5–1.2)a

 

Pure exposure to benzene

2

0.7 (0.1–4.1)a

 

Painters (exposed to benzene)

4

0.8 (0.2–3.3)a

Staines and Cartwright, 1993

Residents of Yorkshire and Trent, UK

 

Benzene, any exposure for any duration

4

2.00 (0.50–8.00)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Other Organic Solvents

Case-Control Study

Nordström et al., 1998

Men in the Swedish Cancer Registry, ever exposed

 

 

White spirit

33

2.0 (1.2–3.4)

 

Acetone

3

1.2 (0.3–4.3)

 

Trichloroethylene

9

1.5 (0.7–3.3)

Unspecified Mixtures of Organic Solvents

Case-Control Studies

Clavel et al., 1998

Male residents of France

 

 

Any exposure

75

0.9 (0.6–1.3)

 

Low intensity

6

1.3 (0.4–4.1)

 

Medium intensity

28

0.9 (0.5–1.5)

 

High intensity

27

1.0 (0.6–1.8)

 

Ever employed as

 

 

Painters

6

1.0 (0.3–3.0)

 

Spray painters

5

2.0 (0.5–8.0)

 

Printers

3

1.4 (0.2–8.5)

 

Launderers and dry cleaners

1

3.0 (0.2–49.2)

Nordström et al., 1998

Men in the Swedish Cancer Registry, ever exposed

 

 

All solvents

51

1.5 (0.99–2.3)

 

Other solvents

5

2.4 (0.8–7.4)

 

Paints

11

4.3 (1.8–10.3)

Clavel et al., 1995

Residents of France, ever exposed

 

 

Men

107

1.2 (0.8–1.7)

 

Women

11

1.1 (0.5–2.4)

Staines and Cartwright, 1993

Residents of Yorkshire and Trent, UK

 

Solvents (any exposure for any duration)

13

1.45 (0.58–3.66)

 

Organic chemicals (past exposure to organic solvents, petroleum products, or benzene)

25

1.30 (0.58–2.93)

aOR is adjusted for smoking and farming.

MYELODYSPLASTIC SYNDROMES

Myelodysplastic syndromes (MDS) are a group of conditions characterized by abnormalities of the bone marrow cells. Because most of the blood cells produced by these abnormal marrow cells are defective, they are usually destroyed before leaving the bone marrow or shortly after entering the bloodstream, and this results in low blood-cell counts. In about 30% of cases, the bone marrow cells develop further abnormalities and eventually develop into acute leukemia. Therefore, the term preleukemia, used in the past to refer to MDS, is not accurate, in that not all patients with MDS develop leukemia.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Description of Case-Control Studies

All case-control studies of MDS reviewed included interviews with study subjects concerning occupational history, some of which yielded information on specific chemical exposures (Table 6.45). Some studies relied on expert review of questionnaire responses (Ciccone et al., 1993; Nisse et al., 2001; West et al., 1995, 2000) to determine exposures; assessments of exposure in other studies were based on self-reports (Goldberg et al., 1990; Ido et al., 1996; Mele et al., 1994; Nagata et al., 1999; Rigolin et al., 1998). Case-control studies of MDS and exposure to organic solvents that had reasonably good assessments of exposure and enough exposed cases include those by Nisse and colleagues (2001) and West and co-workers (1995).

Epidemiologic Studies of Exposure to Organic Solvents

Although several epidemiologic studies were used to examine the association between mixtures of organic solvents and MDS, in only one cohort study of Chinese workers occupationally exposed to benzene (Hayes et al., 1997), was MDS evaluated; however, cases of MDS were combined with cases of ANLL. An RR of 4.1 (95% CI=1.4–11.6) was observed, and the relative risk increased as average exposure to benzene increased (RR=3.2 for constant low-level exposure at under 10 ppm, 7.1 for constant high-level exposure at 25 ppm or higher). This was the first epidemiologic study to demonstrate an association between exposure to benzene and MDS and ANLL.

Earlier case series of benzene-exposed workers noted abnormalities in bone marrow and peripheral blood consistent with MDS specifically (Aksoy and Erdem, 1978; Goguel et al., 1967; Van den Berghe et al., 1979), but the recognition of this disease is relatively recent and has not always been considered in evaluating risks related to exposure to benzene.

Several studies examined the association between exposure to organic solvent mixtures and the risk of MDS. On the basis of qualitative and quantitative assessments of exposure by occupational experts, a study of 204 newly diagnosed cases of MDS and 204 sex- and age-matched controls (Nisse et al., 2001) assessed the association between exposure to solvents and the risk of MDS. Based on 43 exposed cases, the study observed a statistically precise OR of 2.6 (95% CI=1.6–5.4). That increased risk was supported by the other studies reviewed, including a study of 29 cases exposed to organic solvents (Rigolin et al., 1998) that found an OR of 7.11 (95% CI=2.42–20.88). Although those exposed to solvents had an increased risk (OR=7.11. 95% CI=2.42–20.88), those identified as painters, printers, shoemakers, and chemical-industry workers combined did not (OR=0.81, 95% CI=0.33–2.00). The study by Mele and colleagues (1994) found high odds ratios for shoemakers and painters individually (OR=4.3, 95% CI=0.9–21.1 and OR=5.4, 95% CI=0.5–61.0, respectively). The numbers of exposed cases were extremely small, as evidenced by the wide CIs, and the case definition of MDS included those with refractory anemia who had an excess of blasts; this made the health-outcome assessment less precise.

Studies by West and colleagues (2000) and Nagata and colleagues (1999) found increased risks of MDS among residents of the UK (OR=1.8, 95% CI=0.6–6.0) and Japan (OR =1.99, 95% CI=0.97–4.10), respectively. An earlier study in Japan (Ido et al., 1996) also showed increased risks of MDS in men and women exposed to organic solvents occupationally (OR=1.50, 95% CI=0.85–2.64).

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.45 Description of Case-Control Studies of Myelodysplastic Syndromes and Exposure to Organic Solvents

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Goldberg et al., 1990

Cases, age 28–88 years, referred to the hematology-oncology division of the Medical College of Pennsylvania in 1976–1989 with hematologic confirmation; controls selected from the primary-care and cardiology clinics of the same hospital, matched for age, sex, and socioeconomic group

52

52

Solvents

Questionnaire developed by the American Lung Association to assess occupational history and lifetime exposure to solvents (self-reports)

Fisher exact test

None

Ciccone et al., 1993

Cases, age 15–74 years, treated in the Main Hospital of Torino, Italy, newly diagnosed in 1989–1990; hospital controls selected in the same interval, matched for sex, age, and area of residence; population controls randomly selected from residents of Torino, matched on above

19

246

Benzene

Solvents

Interview with standardized questionnaire to assess lifetime occupational history with exposure to solvents determined by industrial hygienist

Logistic regression

Age, area of residence and of birth, smoking

 

Response rates: 91% of cases; 99% of hospital controls; 82% of population controls

 

Mele et al., 1994

Cases, age 15 years or over, identified by hematology departments in three Italian cities in 1986–1990; outpatients without hematological disorders selected as controls

111

(refractory anemia with excess of blasts)

1161

Painting work

Shoemakers

In-hospital interview to assess lifetime behavioral and occupational exposure histories (self-reports)

Logistic regression

Age, sex, education, residence outside study town, other occupations

West et al., 1995

Cases, age 15 years or more, from areas of the UK; controls selected from outpatient clinics and inpatient wards, matched for age, sex, area of residence and hospital, and year of diagnosis (see also West et al., 2000)

400

400

Solvents

In-person interview with questionnaire to assess lifetime occupational and exposure history, duration and intensity of exposure (self-reports)

Matched pairs analysis

Matching variables

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Ido et al., 1996

Cases, age 20–74 years, selected from 32 hospitals in Japan in 1992–1993 with hematologic confirmation; controls selected from outpatient departments of each hospital, matched for age, sex, and hospital

116

116

Organic solvents

Self-administered questionnaire to assess lifetime employment in specific occupations; exposure to solvents was determined from a list of occupations with probable or possible exposure

Conditional logistic regression

Matching variables

Rigolin et al., 1998

Consecutive cases referred to the Institute of Haematology of Ferrara, Italy, in 1990–1996 with hematologic confirmation; controls randomly selected from institute outpatients, matched for sex, date of birth, and geographic area

178

178

Solvents

Occupational exposure: painters, printers, shoemakers, chemical industry

In-person or telephone interview with questionnaire to assess types and duration of exposure (self-reports)

Unadjusted OR

None

Nagata et al., 1999

Cases, age 20–74 years, selected from 28 institutes in Japan in 1995–1996 with hematologic confirmation; controls selected through use of telephone directories, matched for sex and prefecture

111

830

Organic solvents

Self-administered questionnaire to assess occupational exposure to solvents (self-reports)

Conditional logistic regression

Matching variables, age

West et al., 2000

Cases, age 15 years or over, from three specialist regional centers in the UK; controls selected from outpatients, matched for age, sex, and area of residence (see also West et al., 1995)

400

400

Organic chemicals

In-person interview with questionnaire to assess lifetime occupational and exposure history, duration and intensity of exposure

Matched pairs analysis

Matching variables

Nisse et al., 2001

Cases diagnosed in the hematology department of the University Hospital of Lille, France, in 1991–1996 with hematologic confirmation; controls randomly selected from electoral registers, matched for sex and age

204

204

Solvents

In-person interview with questionnaire to assess lifetime occupational history; exposure to solvents was determined by a team of experts

Mantel-Haenszel

Matching variables

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Summary and Conclusion

All but one (Goldberg et al., 1990) of the studies reviewed by the committee and identified below in Table 6.46 found consistently positive odds ratios for the association between MDS and exposure to unspecified mixtures of organic solvents. Given that all the studies reviewed are case-control studies, almost all relied exclusively on self-reported exposures, which may be subject to recall bias. The difficulty in classifying MDS accurately and consistently is another limitation that the committee considered in evaluating the literature.

The committee concludes, from its assessment of the epidemiologic literature, that there is limited/suggestive evidence of an association between chronic exposure to unspecified mixtures of organic solvents and myelodysplastic syndromes.

For exposure to specific solvents, such as benzene, the rarity of the disease and the difficulty in classifying MDS as a separate entity have limited the evaluation of associations in most studies. Additional studies are needed to elucidate and support the association between exposure to benzene and MDS. They should also attempt to separate MDS from other cancers, such as ANLL, so that the relationship between exposure to benzene and MDS can be further understood. Table 6.46 identifies the studies related to MDS and solvent exposure. Unless indicated in the table, the study populations include both men and women.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between chronic exposure to benzene and myelodysplastic syndromes.

TABLE 6.46 Selected Epidemiologic Studies—Myelodysplastic Syndromes and Exposure to Organic Solvents

Reference

Study description

Exposed Cases

Estimated Relative Risk (95% CI)

Benzene

Cohort Study—Incidence

Hayes et al., 1997

Chinese factory workers

 

 

ANLL or MDS

 

 

All unexposed

4

1.0

 

All exposed

28

4.1 (1.4–11.6)

 

Rubber workers

2

6.1

 

Chemical workers

4

4.5

 

Constant exposure, ANLL or MDS

 

 

<10 ppm

10

3.2 (1.0–10.3)

 

10–24 ppm

4

5.1 (1.3–20.6)

 

≥25 ppm

8

7.1 (2.1–23.7)

 

p-trend=0.0003

Unspecified Mixtures of Organic Solvents

Case-Control Studies

Nisse et al., 2001

Residents of northern France

 

 

Solvents, ever exposed

43

2.6 (1.6–5.4)

 

Glue adhesives, ever exposed

11

2.8 (0.8–11.8)

West et al., 2000

Residents of the UK

 

 

Organic chemicals, >50 hours of exposure at moderate intensity

75

1.8 (0.6–6.0)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study description

Exposed Cases

Estimated Relative Risk (95% CI)

Nagata et al., 1999

Residents of Japan

 

 

Organic solvents, occupational exposure

 

 

Males and females

12

1.99 (0.97–4.10)

 

Males

9

1.66 (0.74–3.73)

Rigolin et al., 1998

Residents in Ferrara, Italy

 

 

Solvent, ever exposed

25

7.11 (2.42–20.88)

 

Occupational exposure (painters, printers, shoemakers, chemical industry)

9

0.81 (0.33–2.00)

Ido et al., 1996

Residents of Japan

 

 

Organic solvents, occupational exposure

 

 

Males and females

42

1.50 (0.85–2.64)

 

Males

34

1.28 (0.69–2.36)

 

Females

8

3.50 (0.73–16.8)

West et al., 1995

Residents of the UK

 

 

Solvents, low threshold

38

1.03

 

Solvents, medium threshold

17

0.89

 

Solvents, high threshold

11

1.22

Mele et al., 1994

Residents of Italy

 

 

Shoemakers

3

4.3 (0.9–21.1)a

 

Painters

2

5.4 (0.5–61.0)a

Goldberg et al., 1990

Residents of Pennsylvania

 

 

Solvents, ever exposed

5

0.35 (0.08–1.54)b

aIncludes refractory anemia with excess of blasts. bRisk estimate and 95% CI calculated by the committee using standard methods from the observed and expected numbers presented in the original study.

CHILDHOOD CANCER

Early studies of parental occupation and childhood cancers followed from work by Fabia and Thuy (1974), who found an association between paternal hydrocarbon-related occupations and childhood cancer. A number of studies have since examined potential exposures to hydrocarbons (e.g., DeRoos et al., 2001; Johnson et al., 1987; Van Steensel-Moll, 1985; Zack et al., 1980), and others have focused on parental exposure to solvent mixtures (e.g., Cordier et al., 1997; Smulevich et al., 1999; Shu et al., 1999) as a potential risk factor. Given that the committee was charged with evaluating exposures that occurred during the Gulf War and that soldiers, if they were determined to be pregnant, were removed from the area immediately, the committee did not review studies that considered exposure during pregnancy or after the birth of a child; those studies were not considered directly relevant to the exposure period of interest for this review. Instead, the committee focused its evaluation and conclusions on studies that analyzed preconception exposures. Discussion of the rationale for that exclusion criterion is in Chapter 5 in the childhood cancer section.

Description of Case-Control Studies

Studies of parental occupational exposure and childhood cancer have relied on four approaches to determine exposure: use of occupation or industry as surrogates of exposure, self-reporting of specific exposure, the use of a job-exposure matrix, and review of occupational-

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

history information by an industrial hygienist. Most case-control studies relied on the use of occupational or industrial titles as indexes of exposure. The childhood cancers that were examined include brain tumors (Cordier et al., 1997), leukemia (Lowengart et al., 1987), neuroblastoma (Olshan et al., 1999), and multiple cancers simultaneously (Smulevich et al., 1999). Other case-control studies of childhood cancer used self-reported exposure, including studies of leukemia (Lowengart et al., 1987) or specific subtypes of leukemia—ANLL (Buckley et al., 1989) and ALL (Shu et al., 1999)—and of neuroblastoma (De Roos et al., 2001). A smaller number of studies either used a job-exposure matrix (Cordier et al., 1997; Feingold et al., 1992; Smulevich et al., 1999) or relied on expert review by industrial hygienists (De Roos et al., 2001; Olshan et al., 1999). Table 6.47 describes the study design characteristics of each of the case-control studies reviewed by the committee.

Epidemiologic Studies of Exposure to Organic Solvents and Childhood Leukemia

The literature consisted of studies that analyzed all types of childhood leukemia together and studies that focused on specific leukemia cell types. The committee reported the findings as presented in the literature.

In one cohort study, paternal exposure (Feychting et al., 2001) to solvents and benzene was determined by linking the fathers’ occupational information with a job-exposure matrix; increased RRs of 1.25 for exposure to solvents (95% CI=0.80–1.95) and 1.23 for exposure to benzene (95% CI=0.39–3.85) were found. The exposure period was defined as the period 2–26 months before the child’s birth, which includes both preconception and pregnancy-related exposure. As a result, the study is not completely relevant to this review, given the committee’s focus on preconception exposures.

Two case-control studies examined both maternal and paternal exposures (Lowengart et al., 1987; Smulevich et al., 1999). In a Russian case-control study (Smulevich et al., 1999), maternal occupational exposure to solvents was found to have an OR of 3.1 (95% CI=1.5–6.3). An increased risk of childhood leukemia was also found to be associated with paternal occupational exposure to solvents prior to conception (OR=1.4, 95% CI=0.95–2.1). In a case-control study in Los Angeles County, Lowengart and colleagues (1987) found associations with preconception paternal exposure to chlorinated solvents (OR=2.2, p=0.09), trichloroethylene (OR=2.0, p=0.16), and methyl ethyl ketone (OR=1.7, p=0.24). Mothers of too few cases were occupationally exposed for analysis in this study.

Shu and colleagues (1999) conducted an extensive study of children with ALL and found an association with mothers’ self-reports of exposure to any solvents (OR=1.8, 95% CI=1.3–2.5), to chlorinated solvents (OR=1.8, 95% CI=0.2–20.8), to nonchlorinated organic solvents (OR=2.0, 95% CI=1.0–4.2), to trichloroethylene (OR=1.8, 95% CI=0.6–5.2), to tetrachloroethylene (OR=1.4, 95% CI=0.2–8.6), to toluene (OR=1.5, 95% CI=0.6–3.8), and to paint remover (OR=2.5, 95% CI=1.0–5.9), but not with exposure to benzene (OR=0.7, 95% CI=0.3–1.8) or to methyl ethyl ketone (OR=0.8, 95% CI=0.3–1.9). For preconception paternal exposure, associations were found with exposure to nonchlorinated, organic solvents (OR=1.3, 95% CI=0.8–1.9) and to benzene (OR=1.2, 95% CI=0.8–1.2), but no associations were found with other exposures. Feingold and colleagues (1992) found increased risks of ALL associated with paternal exposure to solvents (OR=1.7, 95% CI=0.4–8.2), to benzene (OR=1.6, 95% CI=0.5–5.8), and to diethylene glycol (OR=1.4, 95% CI=0.4–4.5) in the year before birth.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

TABLE 6.47 Description of Case-Control Studies of Childhood Cancer and Exposure to Organic Solvents

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Lowengart et al., 1987

Cases, age 10 years and under, identified from the Los Angeles County Cancer Surveillance Program in 1980–1984; controls selected from friends of case children or through RDD, matched on age, sex, and race (including native Spanish-speaking Caucasians)

123 leukemia

123

Trichloroethylene

MEK

Chlorinated solvents

Telephone interview with questionnaire to assess parental occupational and exposure history (self-reports)

Matched pairs

Matching variables

 

Participation rate: 79% of cases

 

Buckley et al., 1989

Cases, age under 18 years, selected from the Childrens Cancer Study Group with diagnosis in 1980–1984; controls selected through RDD, matched for date of birth and race

204 ANLL

204

Solvents

Telephone interviews used to assess lifetime occupational history (job titles) and exposure contacts of parents

Conditional logistic regression

Matching variables

Feingold et al., 1992

Cases, age 14 years or under, identified through the Colorado Central Cancer Registry of Denver residents diagnosed in 1976–1983 with almost complete histologic confirmation; controls selected through RDD, matched for age, sex, and telephone exchange area

59 ALL

48 brain

222

Benzene

Diethylene glycol

Solvents

In-person or telephone interview with questionnaire to assess parental occupational history (job-industry titles); job-exposure matrix used to assess exposures

Mantel-Haenszel

Father’s education

 

Participation rates: 70.8% of cases, 79.9% of controls

 

Cordier et al., 1997

Cases, age 15 years or under, identified in centers in Paris, Milan, and Valencia with partial histologic verification (72–87% complete); population controls selected from census information or local health-service records, depending on location, selectively matched for year of birth, sex, and area of residence

251 brain

601

Solvents

In-person interviews to assess parental occupations (job titles) held from 5 years before the child’s birth; job-exposure matrix used to assess exposures

Unconditional logistic regression

Child’s age, sex, exposure to tobacco and ionizing radiation, mother’s age, years of schooling

Olshan et al., 1999

Cases, age under 19 years, selected from the Childrens Cancer Group and Pediatric Oncology Group with diagnosis in 1992–1996; controls selected through RDD, matched for date of birth

504 neuroblastoma

504

Painter (paternal occupation)

Telephone interview to assess parental occupational and exposure history (job-industry titles)

Conditional logistic regression

Maternal race, maternal age, maternal education, household income

 

Participation rates: 73% of cases, 74% of controls

 

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Description of Study Population

Number of Cases

Number of Controls

Relevant Exposure(s)

Determination of Exposure

Analysis

Adjustment for Confounding

Shu et al., 1999

Cases, age under 15 years, diagnosed by a Children’s Cancer Group member or affiliated institution in 1989–1993; controls selected through RDD, matched for age, race, and telephone area code and exchange

1842 ALL

1986

Benzene

Paint remover

Chlorinated solvents

Nonchlorinated solvents

Any solvents

Telephone interview to assess parental occupations (job-industry titles) and exposure history (self-reports)

Conditional and unconditional logistic regression

Matching variables, maternal and paternal education, sex, family income

 

Participation rates: 92% of cases, 76.5% of controls

 

Smulevich et al., 1999

Cases, age 14 years and under, identified from the Moscow Central Cancer Dispensary in 1986-1988; controls randomly selected from records of local pediatric polyclinics in Moscow, matched for age, sex, and residence

593 cancers (number of leukemia cases not specified)

1181 healthy controls

Solvents

In-person interview with questionnaire to assess parental lifetime occupational histories (job titles); exposure to solvents derived from job-exposure matrixes

Conditional logistic regression

Parental alcohol consumption

De Roos et al., 2001

Cases, age under 19 years, selected from the Childrens Cancer Group and Pediatric Oncology Group with diagnosis in 1992–1994; controls selected through RDD, matched for date of birth (see also Olshan et al., 1999)

504 neuroblastoma

504

Trichloroethylene

Benzene

Xylene

Acetone

MEK

Alcohols

Halogenated hydrocarbons

Volatile hydrocarbons

Telephone interview to assess parental occupational and exposure history (self-reports) with coded exposure based on industrial hygienist review

Unconditional logistic regression

Child’s age, maternal race, maternal age, maternal education

 

Participation rates: 73% of cases, 74% of controls (Followup of Olshan et al., 1999)

 

NOTE: ALL=acute lymphocytic leukemia; ANLL=acute nonlymphoblastic leukemia; MEK=methyl ethyl ketone; RDD=random-digit dialing

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Only one study investigated ANLL in conjunction with preconception solvent exposure (Buckley et al., 1989). On the basis of self-reported exposure, a positive association between ANLL and paternal solvent exposure before pregnancy (OR=2.2, p<0.05) was found. No estimates of relative risk for maternal exposure to solvents were presented.

Summary and Conclusion

The studies on childhood leukemia and exposure to organic solvents generally followed similar procedures for ascertaining cases, matching controls, and interviewing parents to obtain relevant information. Studies generally controlled for known confounders, but in most cases risk factors are not well understood. Studies that examined parental occupational exposure often did not differentiate whether the exposure occurred before, during, or after pregnancy. Little information was available on preconception exposure for each cancer type. Exposure measures were based largely on interviews and, thus, were subject to recall bias or random misclassification of exposures; the former tends to artificially increase the odds ratio, and the latter attenuates toward the null value. Many of the studies presented the additional concern that mothers who reported on their husbands’ work exposure further increased the likelihood of misclassification of the fathers’ exposure.

For childhood leukemias combined, several studies showed positive associations with exposure to solvents. Several studies were limited by misclassification bias related to self-reporting of exposure and by the fact that some looked at all childhood leukemias and others focused on specific cell types, such as ALL and ANLL. Given the combination of the limitations of this body of evidence and the consistently positive findings, the committee was unable to reach a consensus conclusion. Some committee members believed that the evidence fulfilled the category of inadequate/insufficient, while others believed it was limited/suggestive of an association. Future studies that address some of the limitations identified above are needed to understand the association between childhood leukemia and exposure to solvents. Table 6.48 identifies all the studies evaluated by the committee.

TABLE 6.48 Selected Epidemiologic Studies—Childhood Leukemia and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Leukemia (all types)

Cohort Study—Incidence

Feychting et al., 2001

Children in Sweden

 

 

Paternal preconception exposure

 

 

Solvents

23

1.25 (0.80–1.95)

 

Benzene

3

1.23 (0.39–3.85)

Case-Control Studies

Smulevich et al., 1999

Children in Moscow

 

 

Occupational exposure before conception

 

 

Maternal exposure to solvents

20

3.1 (1.5–6.3)

 

Paternal exposure to solvents

70

1.4 (0.95–2.1)

Lowengart et al., 1987

Children in Los Angeles County

 

 

Paternal preconception exposure

 

 

Chlorinated solvents

9

2.2 (p=0.09)

 

Trichloroethylene

6

2.0 (p=0.16)

 

Methyl ethyl ketone

5

1.7 (p=0.24)

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Acute Lymphocytic Leukemia

Case-Control Studies

Shu et al., 1999

Children diagnosed through the Children’s Cancer Group

 

 

Maternal preconception exposure, any solvents

93

1.8 (1.3–2.5)

 

Chlorinated solvents

2

1.8 (0.2–20.8)

 

Nonchlorinated organic solvents

22

2.0 (1.0–4.2)

 

Benzene

7

0.7 (0.3–1.8)

 

Trichloroethylene

9

1.8 (0.6–5.2)

 

Tetrachloroethylene

3

1.4 (0.2–8.6)

 

Methyl ethyl ketone

9

0.8 (0.3–1.9)

 

Toluene

5

1.5 (0.6–3.8)

 

Naphtha

2

0.5 (0.1–3.0)

 

Paint remover

16

2.5 (1.0–5.9)

 

Paternal preconception exposure, any solvents

490

1.1 (0.9–1.3)

 

Chlorinated solvents

9

1.0 (0.4–2.5)

 

Nonchlorinated organic solvents

61

1.3 (0.8–1.9)

 

Benzene

74

1.2 (0.8–1.2)

 

Trichloroethylene

100

1.1 (0.8–1.5)

 

Tetrachloroethylene

21

0.8 (0.5–1.5)

 

Methyl Ethyl Ketone

107

1.1 (0.8–1.5)

 

Toluene

82

1.1 (0.8–1.5)

 

Xylene

67

1.2 (0.8–1.8)

 

Naphtha

62

1.2 (0.8–1.7)

 

Paint remover

120

1.0 (0.7–1.3)

Feingold et al., 1992

Children in the Denver, CO, area

 

 

Paternal occupational exposure year before birth

 

 

Solvents

3

1.7 (0.4–8.2)

 

Benzene

9

1.6 (0.5–5.8)

 

Diethylene glycol

7

1.4 (0.4–4.5)

Acute Nonlymphocytic Leukemia

Case-Control Study

Buckley et al., 1989

Children diagnosed through the Children’s Cancer Group

 

 

Paternal preconception exposure, solvents

NA

2.2 (p<0.05)

NOTE: NA=not applicable.

Epidemiologic Studies of Exposure to Organic Solvents and Neuroblastoma

Olshan and colleagues (1999) reported an increased risk of neuroblastoma in children whose fathers were painters (OR=2.1, 95% CI=0.9–4.8); this was the most relevant of the 73 paternal occupations listed for solvent exposure. In a followup study, the investigators used a job-exposure matrix to evaluate maternal and paternal occupational exposure to 65 chemical compounds or broad categories of substances (De Roos et al., 2001). As reviewed by an industrial hygienist, neuroblastoma risk was not markedly increased based on maternal exposures to halogenated hydrocarbons (OR=0.7, 95% CI=0.2–2.1), to volatile hydrocarbons (OR=1.2, 95% CI=0.7–2.1), to acetone (OR=1.1, 95% CI=0.4–2.8), or to alcohols (OR=1.0, 95% CI=0.5–2.1). However, for paternal exposures, risk estimates were higher: volatile hydrocarbons (OR=1.5, 95% CI=1.0–2.1), alcohols (OR=1.8, 95% CI=0.9–3.3), benzene (OR=2.0, 95% CI=0.4–10.3), methyl ethyl ketone (OR=1.4, 95% CI=0.5–3.8), naphtha (OR=1.4, 95% CI=0.4–5.9), and xylene (OR=1.4, 95% CI=0.5–4.3). Paternal exposure to acetone,

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

trichloroethylene, tetrachloroethylene, methylene chloride, and chloroform did not show increased risks.

Summary and Conclusion

The two studies identified were well-conducted. They evaluated a large number of neuroblastoma cases and possible exposures, while considering the possibility of recall bias. However, other corroborating studies are needed to clarify whether an association exists. Table 6.49 identifies the study reviewed by the committee for neuroblastoma.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between either maternal or paternal preconception exposure to solvents under review and neuroblastoma.

TABLE 6.49 Selected Epidemiologic Studies—Childhood Neuroblastoma and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Neuroblastoma

Case-Control Studies

De Roos et al., 2001

Children registered at Children’s Cancer Group or Pediatric Oncology Group hospitals

 

 

Occupational exposure in the 2 years before child’s birtha

 

 

Maternal exposure

 

 

Halogenated hydrocarbons

6

0.7 (0.2–2.1)

 

Volatile hydrocarbons

27

1.2 (0.7–2.1)

 

Acetone

9

1.1 (0.4–2.8)

 

Alcohols

14

1.0 (0.5–2.1)

 

Paternal exposure

 

 

Halogenated hydrocarbons

34

0.9 (0.5–1.5)

 

Trichloroethylene

9

0.9 (0.3–2.5)

 

Volatile hydrocarbons

122

1.5 (1.0–2.1)

 

Acetone

23

0.9 (0.5–1.7)

 

Benzene

5

2.0 (0.4–10.3)

 

Methyl ethyl ketone

12

1.4 (0.5–3.8)

 

Xylene

10

1.4 (0.5–4.3)

 

Tetrachloroethylene

4

0.5 (0.1–1.7)

 

Methylene chloride

4

0.7 (0.2–2.8)

 

Chloroform

3

1.2 (0.2–7.5)

 

Alcohols

49

1.8 (0.9–3.3)

 

Naphtha

11

1.4 (0.4–5.9)

Olshan et al., 1999

Children registered at Children’s Cancer Group or Pediatric Oncology Group hospitals

 

 

Paternal occupation—painter

18

2.1 (0.9–4.8)

aIndustrial hygienist-reviewed exposure information.

Epidemiologic Studies of Exposure to Organic Solvents and Brain Cancer

A European case-control study of brain cancer and parental occupation found that “high” maternal occupational exposure to solvents was strongly associated with an increased risk of brain cancer (OR=2.4, 95% CI=1.2–4.9) and primitive neuroectodermal tumor (PNET) (OR=

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

3.2, 95% CI=1.0–10.3); estimates of risk of astroglial tumors were also increased for “high” levels of exposure to solvents (OR=2.3, 95% CI=0.9–5.8) and other glial tumors were not (OR =0.8, 95% CI=0.1–6.6) (Cordier et al., 1997). Like maternal exposure, high paternal exposure to solvents was associated with increased risk of brain cancer (OR=1.2, 95% CI=0.7–1.9) and astroglial tumors (OR=1.3, 95% CI=0.7–2.3), but not other glial tumors (OR=0.4, 95% CI 0.1–1.4). A second case-control study of childhood brain cancer and paternal occupational exposure showed no association with exposure to benzene (OR=0.7, 95% CI=0.1–3.1) (Feingold et al., 1992). However, an association was found with exposure to solvents (OR=1.2, 95% CI=0.2–8.5) and diethylene glycol (OR=1.3, 95% CI=0.3–5.2).

Summary and Conclusion

On the basis of those two case-control studies, the committee decided that the body of evidence was too small and inconsistent in outcome (brain cancer broadly vs PNET vs astroglial tumors). Table 6.50 identifies the studies reviewed by the committee in drawing its conclusion.

The committee concludes, from its assessment of the epidemiologic literature, that there is inadequate/insufficient evidence to determine whether an association exists between either maternal or paternal preconception exposure to solvents under review and childhood brain cancers.

TABLE 6.50 Selected Epidemiologic Studies—Childhood Brain Cancers and Exposure to Organic Solvents

Reference

Study Population

Exposed Cases

Estimated Relative Risk (95% CI)

Brain Cancer

Case-Control Studies

Cordier et al., 1997

Children in Milan, Paris, and Valencia

 

 

Occupational exposure in the 5 years before birth

 

 

Maternal exposure to solvents

 

 

Medium level, brain cancer

30

1.0 (0.6–1.7)

 

High level, brain cancer

19

2.4 (1.2–4.9)

 

Medium level, PNET

7

1.3 (0.5–3.3)

 

High level, PNET

5

3.2 (1.0–10.3)

 

Medium level, astroglial tumors

15

1.1 (0.6–2.1)

 

High level, astroglial tumors

9

2.3 (0.9–5.8)

 

Medium level, other glial tumors

6

0.9 (0.3–2.6)

 

High level, other glial tumors

1

0.8 (0.1–6.6)

 

Paternal exposure to solvents

 

 

Medium level, brain cancer

40

0.9 (0.6–1.5)

 

High level, brain cancer

37

1.2 (0.7–1.9)

 

Medium level, PNET

16

2.3 (1.1–4.9)

 

High level, PNET

9

1.7 (0.7–4.1)

 

Medium level, astroglial tumors

23

1.1 (0.6–2.0)

 

High level, astroglial tumors

24

1.3 (0.7–2.3)

 

Medium level, other glial tumors

10

1.3 (0.6–3.0)

 

High level, other glial tumors

3

0.4 (0.1–1.4)

Feingold et al., 1992

Children in the Denver, CO, area

 

 

Paternal occupational exposure year before birth

 

 

Benzene

4

0.7 (0.1–3.1)

 

Solvents

2

1.2 (0.2–8.5)

 

Diethylene glycol

7

1.3 (0.3–5.2)

NOTE: PNET=primitive neuroectodermal tumors

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

REFERENCES

Acquavella J, Leet T, Johnson G. 1993. Occupational experience and mortality among a cohort of metal components manufacturing workers. Epidemiology 4(5):428–434.

Aksoy M, Erdem S. 1978. Followup study on the mortality and the development of leukemia in 44 pancytopenic patients with chronic exposure to benzene. Blood 52(2):285–292.

Albin M, Bjork J, Welinder H, Tinnerberg H, Mauritzson N, Johansson B, Billstrom R, Stromberg U, Mikoczy Z, Ahlgren T, Nilsson PG, Mitelman F, Hagmar L. 2000. Acute myeloid leukemia and clonal chromosome aberrations in relation to past exposure to organic solvents. Scandinavian Journal of Work, Environment and Health 26(6):482–491.

Alderson MR, Rattan NS. 1980. Mortality of workers on an isopropyl alcohol plant and two MEK dewaxing plants. British Journal of Industrial Medicine 37(1):85–89.

Anttila A, Pukkala E, Sallmen M, Hernberg S, Hemminki K. 1995. Cancer incidence among Finnish workers exposed to halogenated hydrocarbons. Journal of Occupational and Environmental Medicine 37(7):797–806.

Anttila A, Pukkala E, Riala R, Sallmen M, Hemminki K. 1998. Cancer incidence among Finnish workers exposed to aromatic hydrocarbons. International Archives of Occupational and Environmental Health 71(3):187–193.

Arp EW, Wolf PH, Checkoway H. 1983. Lymphocytic leukemia and exposures to benzene and other solvents in the rubber industry. Journal of Occupational Medicine 25(8):598–602.

Asal NR, Geyer JR, Risser DR, Lee ET, Kadamani S, Cherng N. 1988. Risk factors in renal cell carcinoma. II. Medical history, occupation, multivariate analysis, and conclusions. Cancer Detection and Prevention 13(3–4):263–279.

Aschengrau A, Ozonoff D, Paulu C, Coogan P, Vezina R, Heeren T, Zhang Y. 1993. Cancer risk and tetrachloroethylene-contaminated drinking water in Massachusetts. Archives of Environmental Health 48(5):284–292.

Aschengrau A, Paulu C, Ozonoff D. 1998. Tetrachloroethylene-contaminated drinking water and the risk of breast cancer. Environmental Health Perspectives 106(Suppl 4):947–953.

ATSDR (Agency for Toxic Substances and Disease Registry). 1997a. Toxicological Profile for Benzene. Atlanta, GA: ATSDR.

ATSDR. 1997b. Toxicological Profile for Trichloroethylene. Atlanta, GA: ATSDR.

ATSDR. 1997c. Toxicological Profile for Tetrachloroethylene. Atlanta, GA: ATSDR.

ATSDR. 1997d. Toxicological Profile for Chloroform. Atlanta, GA: ATSDR.

ATSDR. 1998. Toxicological Profile for Phenol. Atlanta, GA: ATSDR.

ATSDR. 2000. Toxicological Profile for Methylene Chloride. Atlanta, GA: ATSDR.

Axelson O, Andersson K, Hogstedt C, Holmberg B, Molina G, de Verdier A. 1978. A cohort study on trichloroethylene exposure and cancer mortality. Journal of Occupational Medicine 20(3):194–196.

Axelson O, Selden A, Andersson K, Hogstedt C. 1994. Updated and expanded Swedish cohort study on trichloroethylene and cancer risk. Journal of Occupational Medicine 36(5):556–562.


Band PR, Le ND, Fang R, Deschamps M, Gallagher RP, Yang P. 2000. Identification of occupational cancer risks in British Columbia. A population-based case-control study of 995 incident breast cancer cases by menopausal status, controlling for confounding factors. Journal of Occupational and Environmental Medicine 42(3):284–310.

Berlin K, Edling C, Persson B, Ahlborg G, Hillert L, Hogstedt B, Lundberg I, Svensson BG, Thiringer G, Orbaek P. 1995. Cancer incidence and mortality of patients with suspected solvent-related disorders. Scandinavian Journal of Work, Environment and Health 21(5):362–367.

Bernard SM, Cartwright RA, Bird CC, Richards ID, Lauder I, Roberts BE. 1984. Aetiologic factors in lymphoid malignancies: A case-control epidemiological study. Leukemia Research 8(4):681–689.

Blair A, Stewart PA, Tolbert PE, Grauman D, Moran FX, Vaught J, Rayner J. 1990. Cancer and other causes of death among a cohort of dry cleaners. British Journal of Industrial Medicine 47(3):162–168.

Blair A, Linos A, Stewart PA, Bermeister LF, Gibson R, Everett G, Schuman L, Cantor KP. 1992. Comments on occupational and environmental factors in the origin of non-Hodgkin’s lymphoma. Cancer Research 52(Suppl 19):5501s–5502s.

Blair A, Hartge P, Stewart PA, McAdams M, Lubin J. 1998. Mortality and cancer incidence of aircraft maintenance workers exposed to trichloroethylene and other organic solvents and chemicals: Extended follow up. Occupational and Environmental Medicine 55(3):161–171.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Boice JD Jr, Marano DE, Fryzek JP, Sadler CJ, McLaughlin JK. 1999. Mortality among aircraft manufacturing workers. Occupational and Environmental Medicine 56(9):581–597.

Bond GG, McLaren EA, Baldwin CL, Cook RR. 1986. An update of mortality among chemical workers exposed to benzene. British Journal of Industrial Medicine 43(10):685–691.

Bond GG, McLaren EA, Sabel FL, Bodner KM, Lipps TE, Cook RR. 1990. Liver and biliary tract cancer among chemical workers. American Journal of Industrial Medicine 18(1):19–24.

Bourguet CC, Checkoway H, Hulka BS. 1987. A case-control study of skin cancer in the tire and rubber manufacturing industry. American Journal of Industrial Medicine 11(4):461–473.

Brown DP, Kaplan SD. 1987. Retrospective cohort mortality study of dry cleaner workers using perchloroethylene. Journal of Occupational Medicine 29(6):535–541.

Brownson RC, Alavanja MC, Chang JC. 1993. Occupational risk factors for lung cancer among nonsmoking women: A case-control study in Missouri (United States). Cancer Causes and Control 4(5):449–454.

Bruckner JV, Warren DA. 2001. Toxic effects of solvents and vapors. In: Klaassen CD, ed. Casarett and Doull’s Toxicology: The Basic Science of Poisons. 6th ed. New York: McGraw-Hill. Pp. 869–916.

Buckley JD, Robison LL, Swotinsky R, Garabrant DH, LeBeau M, Manchester P, Nesbit ME, Odom L, Peters JM, Woods WG, Hammond GD. 1989. Occupational exposures of parents of children with acute nonlymphocytic leukemia: A report from the Childrens Cancer Study Group. Cancer Research 49(14):4030–4037.


Carpenter AV, Flanders WD, Frome EL, Tankersley WG, Fry SA. 1988. Chemical exposures and central nervous system cancers: A case-control study among workers at two nuclear facilities. American Journal of Industrial Medicine 13(3):351–362.

Chiazze L, Ference LD, Wolf PH. 1980. Mortality among automobile assembly workers. I. Spray painters. Journal of Occupational Medicine 22(8):520–526.

Ciccone G, Mirabelli D, Levis A, Gavarotti P, Rege-Cambrin G, Davico L, Vineis P. 1993. Myeloid leukemias and myelodysplastic syndromes: Chemical exposure, histologic subtype and cytogenetics in a case-control study. Cancer Genetics and Cytogenetics 68(2):135–139.

Clavel J, Mandereau L, Cordier S, Le Goaster C, Hemon D, Conso F, Flandrin G. 1995. Hairy cell leukaemia, occupation, and smoking. British Journal of Haematology 91(1):154–161.

Clavel J, Conso F, Limasset JC, Mandereau L, Roche P, Flandrin G, Hemon D. 1996. Hairy cell leukaemia and occupational exposure to benzene. Occupational and Environmental Medicine 53(8):533–539.

Clavel J, Mandereau L, Conso F, Limasset JC, Pourmir I, Flandrin G, Hemon D. 1998. Occupational exposure to solvents and hairy cell leukaemia. Occupational and Environmental Medicine 55(1):59–64.

Cocco P, Figgs L, Dosemeci M, Hayes R, Linet MS, Hsing AW. 1998. Case-control study of occupational exposures and male breast cancer. Occupational and Environmental Medicine 55(9):599–604.

Cordier S, Clavel J, Limasset JC, Boccon-Gibod L, Le MN, Mandereau L, Hemon D. 1993. Occupational resks of bladder cancer in France: A multicentre case-control study. International Journal of Epidemiology 22(3):403–411.

Cordier S, Lefeuvre B, Filippini G, Peris-Bonet R, Farinotti M, Lovicu G, Mandereau L. 1997. Parental occupation, occupational exposure to solvents and polycyclic aromatic hydrocarbons and risk of childhood brain tumors (Italy, France, Spain). Cancer Causes and Control 8(5):688–697.

Costantini AS, Paci E, Miligi L, Buiatti E, Martelli C, Lenzi S. 1989. Cancer mortality among workers in the Tuscan tanning industry. British Journal of Industrial Medicine 46(6):384–388.

Costantini AS, Miligi L, Kriebel D, Ramazzotti V, Rodella S, Scarpi E, Stagnaro E, Tumino R, Fontana A, Masala G, Vigano C, Vindigni C, Crosignani P, Benvenuti A, Vineis P. 2001. A multicenter case-control study in Italy on hematolymphopoietic neoplasms and occupation. Epidemiology 12(1):78–87.

Crump KS. 1994. Risk of benzene-induced leukemia: A sensitivity analysis of the Pliofilm cohort with additional followup and new exposure estimates. Journal of Toxicology and Environmental Health 42(2):219–242.

Crump KS. 1996. Risk of benzene-induced leukemia predicted from the Pliofilm cohort. Environmental Health Perspectives 104(Suppl 6):1437–1441.

Crump K, Allen B. 1984. Quantitative Estimates of Risk of Leukemia From Occupational Exposure to Benzene. Washington, DC: OSHA Docket H 059b Exhibit 152 (Appendix B).


Demers PA, Vaughan TL, Koepsell TD, Lyon JL, Swanson GM, Greenberg RS, Weiss NS. 1993. A case-control study of multiple myeloma and occupation. American Journal of Industrial Medicine 23(4):629–639.

De Roos AJ, Olshan AF, Teschke K, Poole C, Savitz DA, Blatt J, Bondy ML, Pollock BH. 2001. Parental occupational exposures to chemicals and incidence of neuroblastoma in offspring. American Journal of Epidemiology 154(2):106–114.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Dosemeci M, Blair A, Stewart PA, Chandler J, Trush MA. 1991. Mortality among industrial workers exposed to phenol. Epidemiology 2(3):188–193.

Dosemeci M, Cocco P, Chow WH. 1999. Sex differences in risk of renal cell carcinoma and occupational exposures to chlorinated aliphatic hydrocarbons. American Journal of Industrial Medicine 36(1):54–59.

Dumas S, Parent ME, Siemiatycki J, Brisson J. 2000. Rectal cancer and occupational risk factors: A hypothesis-generating, exposure-based case-control study. International Journal of Cancer 87(6):874–879.


Ekstrom AM, Eriksson M, Hansson LE, Lindgren A, Signorello LB, Nyren O, Hardell L. 1999. Occupational exposures and risk of gastric cancer in a population-based case-control study. Cancer Research 59(23):5932–5937.

Engholm G, Englund A. 1982. Cancer incidence and mortality among Swedish painters. In: Mehlman, MA, ed. Advances in Modern Environmental Toxicology. Vol. 2. Princeton Junction, NJ: Senate Press. Pp. 173–185.

Englund A. 1980. Cancer incidence among painters and some allied trades. Journal of Toxicological and Environmental Health 6(5–6):1267–1273.

Eriksson M, Karlsson M. 1992. Occupational and other environmental factors and multiple myeloma: A population based case-control study. British Journal of Industrial Medicine 49(2):95–103.


Fabbro-Peray P, Daures JP, Rossi JF. 2001. Environmental risk factors for non-Hodgkin’s lymphoma: A population-based case-control study in Languedoc-Roussillon, France. Cancer Causes and Control 12(3):201–212.

Fabia J, Thuy TD. 1974. Occupation of father at time of birth of children dying of malignant diseases. British Journal of Preventive and Social Medicine 28(2):98–100.

Feingold L, Savitz DA, John EM. 1992. Use of a job-exposure matrix to evaluate parental occupation and childhood cancer. Cancer Causes and Control 3(2):161–169.

Feychting M, Plato N, Nise G, Ahlbom A. 2001. Paternal occupational exposures and childhood cancer. Environmental Health Perspectives 109(2):193–196.

Flodin U, Andersson L, Anjou C-G, Palm U-B, Vikrot O, Axelson O. 1981. A case-referent study on acute myeloid leukemia, background radiation and exposure to solvents and other agents. Scandinavian Journal of Work, Environment and Health 7(3):169–178.

Fredriksson M, Bengtsson NO, Hardell L, Axelson O. 1989. Colon cancer, physical activity, and occupational exposures. A case-control study. Cancer 63(9):1838–1842.

Friedlander BR, Hearne T, Hall S. 1978. Epidemiologic investigation of employees chronically exposed to methylene chloride. Mortality analysis. Journal of Occupational Medicine 20(10):657–666.

Fritschi L, Siemiatycki J. 1996a. Melanoma and occupation: Results of a case-control study. Occupational and Environmental Medicine 53(3):168–173.

Fritschi L, Siemiatycki J. 1996b. Lymphoma, myeloma and occupation: Results of a case-control study. International Journal of Cancer 67(4):498–503.

Fu H, Demers PA, Costantini AS, Winter P, Colin D, Kogevinas M, Boffetta P. 1996. Cancer mortality among shoe manufacturing workers: An analysis of two cohorts. Occupational and Environmental Medicine 53(6):394–398.


Garabrant DH, Held J, Langholz B, Bernstein L. 1988. Mortality of aircraft manufacturing workers in southern California. American Journal of Industrial Medicine 13(6):683–693.

Gérin M, Siemiatycki J, Desy M, Krewski D. 1998. Associations between several sites of cancer and occupational exposure to benzene, toluene, xylene, and styrene: Results of a case-control study in Montreal. American Journal of Industrial Medicine 34(2):144–156.

Gibbs GW, Amsel J, Soden K. 1996. A cohort mortality study of cellulose triacetate-fiber workers exposed to methylene chloride. Journal of Occupational and Environmental Medicine 38(7):693–697.

Goguel A, Cavigneaux A, Bernard J. 1967. Benzene leukemias in the Paris region between 1950 and 1965 (study of 50 cases). Nouvelle Revue Francaise d’Hematologie 7(4):465–480.

Goldberg H, Lusk E, Moore J, Nowell PC, Besa EC. 1990. Survey of exposure to genotoxic agents in primary myelodysplastic syndrome: Correlation with chromosome patterns and data on patients without hematological disease. Cancer Research 50(21):6876–6881.

Goldberg MS, Parent ME, Siemiatycki J, Desy M, Nadon L, Richardson L, Lakhani R, Latreille B, Valois MF. 2001. A case-control study of the relationship between the risk of colon cancer in men and exposures to occupational agents. American Journal of Industrial Medicine 39(6):531–546.

Greenland S, Salvan A, Wegman DH, Hallock MF, Smith TJ. 1994. A case-control study of cancer mortality at a transformer-assembly facility. International Archives of Occupational and Environmental Health 66(1):49–54.

Guberan E, Usel M, Raymond L, Tissot R, Sweetnam PM. 1989. Disability, mortality, and incidence of cancer among Geneva painters and electricians: A historical prospective study. British Journal of Industrial Medicine 46(1):16–23.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Hansen J. 1999. Breast cancer risk among relatively young women employed in solvent-using industries. American Journal of Industrial Medicine 36(1):43–47.

Hansen J. 2000. Elevated risk for male breast cancer after occupational exposure to gasoline and vehicular combustion products. American Journal of Industrial Medicine 37(4):349–352.

Hansen J, Raaschou-Nielsen O, Christensen JM, Johansen I, McLaughlin JK, Lipworth L, Blot WJ, Olsen JH. 2001. Cancer incidence among Danish workers exposed to trichloroethylene. Journal of Occupational and Environmental Medicine 43(2):133–139.

Hardell L, Bengtsson NO. 1983. Epidemiological study of socioeconomic factors and clinical findings in Hodgkin’s disease, and reanalysis of previous data regarding chemical exposure. British Journal of Cancer 48(2):217–225.

Hardell L, Eriksson M, Lenner P, Lundgren E. 1981. Malignant lymphoma and exposure to chemicals, especially organic solvents, chlorophenols and phenoxy acids: A case-control study. British Journal of Cancer 43(2):169–176.

Hardell L, Johansson B, Axelson O. 1982. Epidemiological study of nasal and nasopharyngeal cancer and their relation to phenoxy acid or chlorophenol exposure. American Journal of Industrial Medicine 3(3):247–257.

Hardell L, Bengtsson NO, Jonsson U, Eriksson S, Larsson LG. 1984. Aetiological aspects on primary liver cancer with special regard to alcohol, organic solvents and acute intermittent porphyria—An epidemiological investigation. British Journal of Cancer 50(3):389–397.

Hardell L, Eriksson M, Degerman A. 1994. Exposure to phenoxyacetic acids, chlorophenols, or organic solvents in relation to histopathology, stage, and anatomical localization of non- Hodgkin’s lymphoma. Cancer Research 54(9):2386–2389.

Harrington JM, Whitby H, Gray CN, Reid FJ, Aw TC, Waterhouse JA. 1989. Renal disease and occupational exposure to organic solvents: A case-referent approach. British Journal of Industrial Medicine 46(9):643–650.

Hayes RB, Yin SN, Dosemeci M, Li GL, Wacholder S, Chow WH, Rothman N, Wang YZ, Dai TR, Chao XJ, Jiang ZL, Ye PZ, Zhao HB, Kou QR, Zhang WY, Meng JF, Zho JS, Lin XF, Ding CY, Li CY, Zhang ZN, Li DG, Travis LB, Blot WJ, Linet MS. 1996. Mortality among benzene-exposed workers in China. Environmental Health Perspectives 104(Suppl 6):1349–1352.

Hayes RB, Yin SN, Dosemeci M, Li GL, Wacholder S, Travis LB, Li CY, Rothman N, Hoover RN, Linet MS. 1997. Benzene and the dose-related incidence of hematologic neoplasms in China. Chinese Academy of Preventive Medicine—National Cancer Institute Benzene Study Group. Journal of the National Cancer Institute 89(14):1065–1071.

Hearne FT, Friedlander BR. 1981. Followup of methylene chloride study. Journal of Occupational Medicine 23(10):660.

Hearne FT, Pifer JW. 1999. Mortality study of two overlapping cohorts of photographic film base manufacturing employees exposed to methylene chloride. Journal of Occupational and Environmental Medicine 41(12):1154–1169.

Hearne FT, Grose F, Pifer JW, Friedlander BR, Raleigh RL. 1987. Methylene chloride mortality study: Dose-response characterization and animal model comparison. Journal of Occupational Medicine 29(3):217–228.

Hearne FT, Pifer JW, Grose F. 1990. Absence of adverse mortality effects in workers exposed to methylene chloride: An update. Journal of Occupational Medicine 32(3):234–240.

Heineman EF, Cocco P, Gomez MR, Dosemeci M, Stewart PA, Hayes RB, Zahm SH, Thomas TL, Blair A. 1994. Occupational exposure to chlorinated aliphatic hydrocarbons and risk of astrocytic brain cancer. American Journal of Industrial Medicine 26(2):155–169.

Heineman EF, Gao Y-T, Dosemeci M, McLaughlin JK. 1995. Occupational risk factors for brain tumors among women in Shanghai, China. Journal of Occupational and Environmental Medicine 37(3):288–293.

Heinemann K, Willich SN, Heinemann LAJ, DoMinh T, Mohner M, Heuchert GE. 2000. Occupational exposure and liver cancer in women: Results of the Multicentre International Liver Tumour Study (MILTS). Occupational Medicine 50(6):422–429.

Henschler D, Vamvakas S, Lammert M, Dekant W, Kraus B, Thomas B, Ulm K. 1995. Increased incidence of renal cell tumors in a cohort of cardboard workers exposed to trichloroethene. Archives of Toxicology 69(5):291–299.

Hernberg S, Kauppinen T, Riala R, Korkala ML, Asikainen U. 1988. Increased risk for primary liver cancer among women exposed to solvents. Scandinavian Journal of Work, Environment and Health 14(6):356–365.

Holly EA, Lele C, Bracci P. 1997. Non-hodgkin’s lymphoma in homosexual men in the San Francisco bay area: Occupational, chemical, and environmental exposures. Journal of Acquired Immune Deficiency Syndromes and Human Retrovirology 15(3):223–231.

Hunting KL, Longbottom H, Kalavar SS, Stern F, Schwartz E, Welch LS. 1995. Haematopoietic cancer mortality among vehicle mechanics. Occupational and Environmental Medicine 52(10):673–678.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

IARC (International Agency for Research on Cancer). 1987. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. Overall Evaluations of Carcinogenicity: An Updating of IARC Monographs, Volumes 1 to 42. Suppl. 7. Lyon, France: IARC.

IARC. 1989. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. Some Organic Solvents, Resin Monomers and Related Compounds, Pigments and Occupational Exposures in Paint Manufacture and Painting. Vol. 47. Lyon, France: IARC.

IARC. 1995. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. Dry-cleaning, Some Chlorinated Solvents and Other Industrial Chemicals. Vol. 63. Lyon, France: IARC.

IARC. 1999. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. Some Chemicals that Cause Tumours of the Kidney or Urinary Bladder in Rodents, and Some Other Substances. Vol. 73. Lyon, France: IARC.

Ido M, Nagata C, Kawakami N, Shimizu H, Yoshida Y, Nomura T, Mizoguchi H. 1996. A case-control study of myelodysplastic syndromes among Japanese men and women. Leukemia Research 20(9):727–731.

Infante PF, Rinsky RA, Wagoner JK, Young RJ. 1977. Leukaemia in benzene workers. Lancet 2(8028):76–78.

Ireland B, Collins JJ, Buckley CF, Riordan SG. 1997. Cancer mortality among workers with benzene exposure. Epidemiology 8(3):318–320.

Irons RD. 1992. Benzene and other hemotoxins. In: Sullivan JB, Krieger GR, Eds. Hazardous Materials Toxicology: Clinical Principles of Environmental Health. Baltimore: Williams & Wilkins. Pp. 718–731.


Jensen OM, Wahrendorf J, Knudsen JB, Sorensen BL. 1987. The Copenhagen case-referent study on bladder cancer. Risks among drivers, painters and certain other occupations. Scandinavian Journal of Work, Environment and Health 13(2):129–134.

Jensen OM, Knudsen JB, McLaughlin JK, Sorensen BL. 1988. The Copenhagen case-control study of renal pelvis and ureter cancer: Role of smoking and occupational exposures. International Journal of Cancer 41(4):557–561.

Ji BT, Silverman DT, Dosemeci M, Dai Q, Gao YT, Blair A. 1999. Occupation and pancreatic cancer risk in Shanghai, China. American Journal of Industrial Medicine 35(1):76–81.

Johnson CC, Annegers JF, Frankowski RF, Spitz MR, Buffler PA. 1987. Childhood nervous system tumors—an evaluation of the association with paternal occupational exposure to hydrocarbons. American Journal of Epidemiology 126(4):605–613.


Kaerlev L, Teglbjaerg PS, Sabroe S, Kolstad HA, Ahrens W, Eriksson M, Gonzalez AL, Guenel P, Hardell L, Launoy G, Merler E, Merletti F, Suarez-Varela MM, Stang A. 2000. Occupation and small bowel adenocarcinoma: A European case-control study. Occupational and Environmental Medicine 57(11):760–766.

Kauppinen TP, Partanen TJ, Hernberg SG, Nickels JI, Luukkonen RA, Hakulinen TR, Pukkala EI. 1993. Chemical exposures and respiratory cancer among Finnish woodworkers. British Journal of Industrial Medicine 50(2):143–148.

Kauppinen T, Partanen T, Degerth R, Ojajarvi A. 1995. Pancreatic cancer and occupational exposures. Epidemiology 6(5):498–502.


Lagorio S, Forastiere F, Iavarone I, Rapiti E, Vanacore N, Perucci CA, Carere A. 1994. Mortality of filling station attendants. Scandinavian Journal of Work, Environment and Health 20(5):331–338.

Lanes SF, Cohen A, Rothman KJ, Dreyer NA, Soden KJ. 1990. Mortality of cellulose fiber production workers. Scandinavian Journal of Work, Environment and Health 16(4):247–251.

Lanes SF, Rothman KJ, Dreyer NA, Soden KJ. 1993. Mortality update of cellulose fiber production workers. Scandinavian Journal of Work, Environment and Health 19(6):426–428.

La Vecchia C, Negri E, D’Avanzo B, Franceschi S. 1990. Occupation and the risk of bladder cancer. International Journal of Epidemiology 19(2):264–268.

Lazarov D, Waldron HA, Pejin D. 2000. Acute myeloid leukaemia and exposure to organic solvents—A case-control study. European Journal of Epidemiology 16(3):295–301.

Leffingwell SS, Waxweiler R, Alexander V, Ludwig HR, Halperin W. 1983. Case-control study of gliomas of the brain among workers employed by a Texas City, Texas chemical plant, USA. Neuroepidemiology 2(3–4):179–195.

Leon DA. 1994. Mortality in the British printing industry: A historical cohort study of trade union members in Manchester. Occupational and Environmental Medicine 51(2):79–86.

Li GL, Linet MS, Hayes RB, Yin SN, Dosemeci M, Wang YZ, Chow WH, Jiang ZL, Wacholder S, Zhang WU, et al. 1994. Gender differences in hematopoietic and lymphoproliferative disorders and other cancer risks by major occupational group among workers exposed to benzene in China. Journal of Occupational Medicine 36(8):875–881.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Logsdon JE, Loke RA. 1996. Isopropyl alcohol. In: Kirk RE, Kroschwitz JI, Howe-Grant M, eds. Encyclopedia of Chemical Technology. Vol. 20. 4th ed. New York: John Wiley & Sons.

Lowengart RA, Peters JM, Cicioni C, Buckley J, Bernstein L, Preston-Martin S, Rappaport E. 1987. Childhood leukemia and parents’ occupational and home exposures. Journal of the National Cancer Institute 79(1):39–46.

Lundberg I. 1986. Mortality and cancer incidence among Swedish paint industry workers with long-term exposure to organic solvents. Scandinavian Journal of Work, Environment and Health 12(2):108–13.

Lundberg I, Milatou-Smith R. 1998. Mortality and cancer incidence among Swedish paint industry workers with long-term exposure to organic solvents. Scandinavian Journal of Work, Environment and Health 24(4):270–275.

Lynge E, Thygesen L. 1990. Primary liver cancer among women in laundry and dry-cleaning work in Denmark. Scandinavian Journal of Work, Environment and Health 16(2):108–112.

Lynge E, Rix BA, Villadsen E, Andersen I, Hink M, Olsen E, Moller UL, Silfverberg E. 1995. Cancer in printing workers in Denmark. Occupational and Environmental Medicine 52(11):738–744.

Lynge E, Andersen A, Nilsson R, Barlow L, Pukkala E, Nordlinder R, Boffetta P, Grandjean P, Heikkila P, Horte LG, Jakobsson R, Lundberg I, Moen B, Partanen T, Riise T. 1997. Risk of cancer and exposure to gasoline vapors. American Journal of Epidemiology 145(5):449–458.


Malker HS, Gemne G. 1987. A register-epidemiology study on cancer among Swedish printing industry workers. Archives of Environmental Health 42(2):73–82.

Malone KE, Koepsell TD, Daling JR, Weiss NS, Morris PD, Taylor JW, Swanson GM, Lyon JL. 1989. Chronic lymphocytic leukemia in relation to chemical exposures. American Journal of Epidemiology 130(6):1152–1158.

Mandel JS, McLaughlin JK, Schlehofer B, Mellemgaard A, Helmert U, Lindblad P, McCredie M, Adami HO. 1995. International renal-cell cancer study. IV. Occupation. International Journal of Cancer 61(5):601–605.

Matanoski GM, Stockwell HG, Diamond EL, Haring-Sweeney M, Joffe RD, Mele LM, Johnson ML. 1986. A cohort mortality study of painters and allied tradesmen. Scandinavian Journal of Work, Environment and Health 12(1):16–21.

McCredie M, Stewart JH. 1993. Risk factors for kidney cancer in New South Wales. IV. Occupation. British Journal of Industrial Medicine 50(4):349–354.

McMichael AJ, Spirtas R, Kupper LL, Gamble JF. 1975. Solvent exposure and leukemia among rubber workers: An epidemiologic study. Journal of Occupational Medicine 17(4):234–239.

McMichael AJ, Andjelkovic DA, Tyroler HA. 1976. Cancer mortality among rubber workers: An epidemiologic study. Annals of the New York Academy of Sciences 271:125–137.

Mele A, Szklo M, Visani G, Stazi MA, Castelli G, Pasquini P, Mandelli F. 1994. Hair dye use and other risk factors for leukemia and pre-leukemia: A case-control study. Italian Leukemia Study Group. American Journal of Epidemiology 139(6):609–619.

Mellemgaard A, Engholm G, McLaughlin JK, Olsen JH. 1994. Occupational risk factors for renal-cell carcinoma in Denmark. Scandinavian Journal of Work, Environment and Health 20(3):160–165.

Morgan RW, Kaplan SD, Gaffey WR. 1981. A general mortality study of production workers in the paint and coatings manufacturing industry. A preliminary report. Journal of Occupational Medicine 23(1):13–21.

Morgan RW, Kelsh MA, Zhao K, Heringer S. 1998. Mortality of aerospace workers exposed to trichloroethylene. Epidemiology 9(4):424–431.

Morris PD, Koepsell TD, Daling JR, Taylor JW, Lyon JL, Swanson GM, Child M, Weiss NS. 1986. Toxic substance exposure and multiple myeloma: A case-control study. Journal of the National Cancer Institute 76(6):987–994.

Morrison AS, Ahlbom A, Verhoek WG, Aoki K, Leck I, Ohno Y, Obata K. 1985. Occupation and bladder cancer in Boston, USA, Manchester, UK, and Nagoya, Japan. Journal of Epidemiology and Community Health 39(4):294–300.


Nagata C, Shimizu H, Hirashima K, Kakishita E, Fujimura K, Niho Y, Karasawa M, Oguma S, Yoshida Y, Mizoguchi H. 1999. Hair dye use and occupational exposure to organic solvents as risk factors for myelodysplastic syndrome. Leukemia Research 23(1):57–62.

NCI (National Cancer Institute). 2002. What You Need to Know About Cancer of the Cervix: Information About Detection, Symptoms, Diagnosis, and Treatment of Cervical Cancer. Available: http://www.nci.nih.gov/cancer_information/cancer_type/cervical [accessed May 2002].

Nielsen H, Henriksen L, Olsen JH. 1996. Malignant melanoma among lithographers. Scandinavian Journal of Work, Environment and Health 22(2):108–111.

NIOSH (National Institute for Occupational Safety and Health). 1997. NIOSH Pocket Guide to Chemical Hazards. Cincinnati, OH: NIOSH

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Nisse C, Haguenoer JM, Grandbastien B, Preudhomme C, Fontaine B, Brillet JM, Lejeune R, Fenaux P. 2001. Occupational and environmental risk factors of the myelodysplastic syndromes in the North of France. British Journal of Haematology 112(4):927–935.

Nordström M, Hardell L, Magnuson A, Hagberg H, Rask-Andersen A. 1998. Occupational exposures, animal exposure and smoking as risk factors for hairy cell leukaemia evaluated in a case-control study. British Journal of Cancer 77(11):2048–2052.

NTP (National Toxicology Program). 2001. 9th Report of Carcinogens. Research Triangle Park, NC: NTP.


Olshan AF, De Roos AJ, Teschke K, Neglia JP, Strain DO, Pollock BH, Castleberry RP. 1999. Neuroblastoma and parental occupation. Cancer Causes and Control 10(6):539–549.

Olsson H, Brandt L. 1988. Risk of non-Hodgkin’s lymphoma among men occupationally exposed to organic solvents. Scandinavian Journal of Work, Environment and Health 14(4):246–251.

Ott MG, Skory LK, Holder BB, Bronson JM, Williams PR. 1983. Health evaluation of employees occupationally exposed to methylene chloride. Mortality. Scandinavian Journal of Work, Environment and Health 9(Suppl 1):8–16.


Paci E, Buiatti E, Seniori CA, Miligi L, Pucci N, Scarpelli A, Petrioli G, Simonato L, Winkelmann R, Kaldor JM. 1989. Aplastic anemia, leukemia and other cancer mortality in a cohort of shoe workers exposed to benzene. Scandinavian Journal of Work, Environment and Health 15(5):313–318.

Paganini-Hill A, Glazer E, Henderson BE, Ross RK. 1980. Cause-specific mortality among newspaper web pressmen. Journal of Occupational Medicine 22(8):542–544.

Pannett B, Coggon D, Acheson ED. 1985. A job-exposure matrix for use in population based studies in England and Wales. British Journal of Industrial Medicine 42(11):777–783.

Parent ME, Siemiatycki J, Fritschi L. 2000. Workplace exposures and oesophageal cancer. Occupational and Environmental Medicine 57(5):325–334.

Parkes HG, Veys CA, Waterhouse JA, Peters A. 1982. Cancer mortality in the British rubber industry. British Journal of Industrial Medicine 39(3):209–220.

Partanen T, Heikkila P, Hernberg S, Kauppinen T, Moneta G, Ojajarvi A. 1991. Renal cell cancer and occupational exposure to chemical agents. Scandinavian Journal of Work, Environment and Health 17(4):231–239.

Partanen T, Kauppinen T, Luukkonen R, Hakulinen T, Pukkala E. 1993. Malignant lymphomas and leukemias, and exposures in the wood industry: An industry-based case-referent study. International Archives of Occupational and Environmental Health 64(8):593–596.

Paulu C, Aschengrau A, Ozonoff D. 1999. Tetrachloroethylene-contaminated drinking water in Massachusetts and the risk of colon-rectum, lung, and other cancers. Environmental Health Perspectives 107(4):265–271.

Paustenbach DJ, Price PS, Ollison W, Blank C, Jernigan JD, Bass RD, Peterson HD. 1992. Reevaluation of benzene exposure for the Pliofilm (rubberworker) cohort (1936–1976). Journal of Toxicology and Environmental Health 36(3):177–231.

Paxton MB, Chinchilli VM, Brett SM, Rodricks JV. 1994a. Leukemia risk associated with benzene exposure in the Pliofilm cohort: I. Mortality update and exposure distribution. Risk Analysis 14(2):147–154.

Paxton MB, Chinchilli VM, Brett SM, Rodricks JV. 1994b. Leukemia risk associated with benzene exposure in the Pliofilm cohort. II. Risk estimates. Risk Analysis 14(2):155–161.

Paxton MB. 1996. Leukemia risk associated with benzene exposure in the Pliofilm cohort. Environmental Health Perspectives 104(Suppl 6):1431–1436.

Persson B, Fredriksson M. 1999. Some risk factors for non-Hodgkin’s lymphoma. International Journal of Occupational Medicine and Environmental Health 12(2):135–142.

Persson B, Dahlander A-M, Fredriksson M, Noorlind BH, Ohlson C-G, Axelson O. 1989. Malignant lymphomas and occupational exposures. British Journal of Industrial Medicine 46(8):516–520.

Persson B, Fredriksson M, Olsen K, Boeryd B, Axelson O. 1993. Some occupational exposures as risk factors for malignant lymphomas. Cancer 72(5):1773–1778.

Pesch B, Haerting J, Ranft U, Klimpel A, Oelschlagel B, Schill W. 2000a. Occupational risk factors for urothelial carcinoma: Agent-specific results from a case-control study in Germany. MURC Study Group. Multicenter Urothelial and Renal Cancer. International Journal of Epidemiology 29(2):238–247.

Pesch B, Haerting J, Ranft U, Klimpel A, Oelschlagel B, Schill W. 2000b. Occupational risk factors for renal cell carcinoma: Agent-specific results from a case-control study in Germany. MURC Study Group. Multicenter Urothelial and Renal Cancer study. International Journal of Epidemiology 29(6):1014–1024.

Petralia SA, Vena JE, Freudenheim JL, Dosemeci M, Michalek A, Goldberg MS, Brasure J, Graham S. 1999. Risk of premenopausal breast cancer in association with occupational exposure to polycyclic aromatic hydrocarbons and benzene. Scandinavian Journal of Work, Environment and Health 25(3):215–221.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Pippard EC, Acheson ED. 1985. The mortality of boot and shoe makers, with special reference to cancer. Scandinavian Journal of Work, Environment and Health 11(4):249–255.

Pohlabeln H, Boffetta P, Ahrens W, Merletti F, Agudo A, Benhamou E, Benhamou S, Bruske-Hohlfeld K, Ferro G, Fortes C, Kreuzer M, Mendes A, Nyberg F, Pershagen G, Saracci R, Schmid G, Siemiatycki J, Simonato L, Whitley E, Wichmann HE, Winck C, Zambon P, Jockel KH. 2000. Occuptional risks for lung cancer among nonsmokers. Epidemiology 11(5):532–538.

Poole C, Dreyer NA, Satterfleld MH, Levin L, Rothman KJ. 1993. Kidney cancer and hydrocarbon exposures among petroleum refinery workers. Environmental Health Perspectives 101(Suppl 6):53–62.


Richardson S, Zittoun R, Bastuji-Garin S, Lasserre V, Guihenneuc C, Cadiou M, Viguie F, Laffont-Faust I. 1992. Occupational risk factors for acute leukaemia: A case-control study. International Journal of Epidemiology 21(6):1063–1073.

Rigolin GM, Cuneo A, Roberti MG, Bardi A, Bigoni R, Piva N, Minotto C, Agostini P, De Angeli C, Del Senno L, Spanedda R, Castoldi G. 1998. Exposure to myelotoxic agents and myelodysplasia: Case-control study and correlation with clinicobiological findings. British Journal of Haematology 103(1):189–197.

Rinsky RA, Young RJ, Smith AB. 1981. Leukemia in benzene workers. American Journal of Industrial Medicine 2(3):217–245.

Rinsky RA, Smith AB, Hornung R, Filloon TG, Young RJ, Okun AH, Landrigan PJ. 1987. Benzene and leukemia. An epidemiologic risk assessment. New England Journal of Medicine 316(17):1044–1050.

Risch HA, Burch JD, Miller AB, Hill GB, Steele R, Howe GR. 1988. Occupational factors and the incidence of cancer of the bladder in Canada. British Journal of Industrial Medicine 45(6):361–367.

Ritz B. 1999. Cancer mortality among workers exposed to chemicals during uranium processing. Journal of Occupational and Environmental Medicine 41(7):556–566.

Rodvall Y, Ahlbom A, Spannare B, Nise G. 1996. Glioma and occupational exposure in Sweden, a case-control study. Occupational and Environmental Medicine 53(8):526–537.

Ruder AM, Ward EM, Brown DP. 1994. Cancer mortality in female and male dry-cleaning workers. Journal of Occupational Medicine 36(8):867–874.

Ruder AM, Ward EM, Brown DP. 2001. Mortality in dry-cleaning workers: An update. American Journal of Industrial Medicine 39(2):121–132.

Rushton L, Alderson MR. 1981. A case-control study to investigate the association between exposure to benzene and deaths from leukaemia in oil refinery workers. British Journal of Cancer 43(1):77–84.

Rushton L, Romaniuk H. 1997. A case-control study to investigate the risk of leukaemia associated with exposure to benzene in petroleum marketing and distribution workers in the United Kingdom. Occupational and Environmental Medicine 54(3):152–166.


Scherr PA, Hutchison GB, Neiman RS. 1992. Non-Hodgkin’s lymphoma and occupational exposure. Cancer Research 52(Suppl 19):5503–5509.

Schlehofer B, Heuer C, Blettner M, Niehoff D, Wahrendorf J. 1995. Occupation, smoking and demographic factors, and renal cell carcinoma in Germany. International Journal of Epidemiology 24(1):51–57.

Schnatter AR, Armstrong TW, Nicolich MJ, Thompson FS, Katz AM, Huebner WW, Pearlman ED. 1996a. Lymphohaematopoietic malignancies and quantitative estimates of exposure to benzene in Canadian petroleum distribution workers. Occupational and Environmental Medicine 53(11):773–781.

Schnatter AR, Armstrong TW, Thompson LS, Nicolich MJ, Katz AM, Huebner WW, Pearlman ED. 1996b. The relationship between low-level benzene exposure and leukemia in Canadian petroleum distribution workers. Environmental Health Perspectives 104(Suppl 6):1375–1379.

Schoenberg JB, Stemhagen A, Mogielnicki AP, Altman R, Abe T, Mason TJ. 1984. Case-control study of bladder cancer in New Jersey. I. Occupational exposures in white males. Journal of the National Cancer Institute 72(5):973–981.

Schumacher MC, Delzell E. 1988. A death-certificate case-control study of non-Hodgkin’s lymphoma and occupation in men in North Carolina. American Journal of Industrial Medicine 13(3):317–330.

Serraino D, Franceschi S, La Vecchia C, Carbone A. 1992. Occupation and soft-tissue sarcoma in northeastern Italy. Cancer Causes and Control 3(1):25–30.

Shannon HS, Haines T, Bernholz C, Julian JA, Verma DK, Jamieson E, Walsh C. 1988. Cancer morbidity in lamp manufacturing workers. American Journal of Industrial Medicine 14(3):281–290.

Sharpe CR, Rochon JE, Adam JM, Suissa S. 1989. Case-control study of hydrocarbon exposures in patients with renal cell carcinoma. Canadian Medical Association Journal 140(11):1309–1318.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Shu XO, Stewart P, Wen WQ, Han D, Potter JD, Buckley JD, Heineman E, Robison LL. 1999. Parental occupational exposure to hydrocarbons and risk of acute lymphocytic leukemia in offspring. Cancer Epidemiology, Biomarkers and Prevention 8(9):783–791.

Silverman DT, Levin LI, Hoover RN, Hartge P. 1989a. Occupational risks of bladder cancer in the United States: I. White men. Journal of the National Cancer Institute 81(19):1472–1480.

Silverman DT, Levin LI, Hoover RN. 1989b. Occupational risks of bladder cancer in the United States: II. Nonwhite men. Journal of the National Cancer Institute 81(19):1480–1483.

Smith EM, Miller ER, Woolson RF, Brown CK. 1985. Bladder cancer risk among laundry workers, dry cleaners, and others in chemically-related occupations. Journal of Occupational Medicine 27(4):295–297.

Smulevich VB, Solionova LG, Belyakova SV. 1999. Parental occupation and other factors and cancer risk in children: II. Occupational factors. International Journal of Cancer 83(6):718–722.

Sorahan T, Cathcart M. 1989. Lung cancer mortality among workers in a factory manufacturing chlorinated toluenes: 1961–84. British Journal of Industrial Medicine 46(6):425–427.

Sorahan T, Parkes HG, Veys CA, Waterhouse JA. 1986. Cancer mortality in the British rubber industry: 1946–1980. British Journal of Industrial Medicine 43(6):363–373.

Sorahan T, Parkes HG, Veys CA, Waterhouse JA, Straughan JK, Nutt A. 1989. Mortality in the British rubber industry 1946–85. British Journal of Industrial Medicine 46(1):1–10.

Spirtas R, Stewart PA, Lee JS, Marano DE, Forbes CD, Grauman DJ, Pettigrew HM, Blair A, Hoover RN, Cohen JL. 1991. Retrospective cohort mortality study of workers at an aircraft maintenance facility. I. Epidemiological results. British Journal of Industrial Medicine 48(8):515–530.

Staines A, Cartwright RA. 1993. Hairy cell leukaemia: Descriptive epidemiology and a case-control study. British Journal of Haematology 85(4):714–717.

Steenland K, Palu S. 1999. Cohort mortality study of 57,000 painters and other union members: A 15 year update. Occupational and Environmental Medicine 56(5):315–321.

Stemhagen A, Slade J, Altman R, Bill J. 1983. Occupational risk factors and liver cancer. A retrospective case-control study of primary liver cancer in New Jersey. American Journal of Epidemiology 117(4):443–454.

Stern FB, Waxweiler RA, Beaumont JJ, Lee ST, Rinsky RA, Zumwalde RD, Halperin WE, Bierbaum PJ, Landrigan PJ, Murray WE. 1986. A case-control study of leukemia at a naval nuclear shipyard. American Journal of Epidemiology 123(6):980–992.

Stewart PA, Lee JS, Marano DE, Spirtas R, Forbes CD, Blair A. 1991. Retrospective cohort mortality study of workers at an aircraft maintenance facility. II. Exposures and their assessment. British Journal of Industrial Medicine 48(8):531–537.

Stockwell HG, Matanoski GM. 1985. A case-control study of lung cancer in painters. Journal of Occupational Medicine 27(2):125–126.

Svensson BG, Nise G, Englander V, Attewell R, Skerfving S, Moller T. 1990. Deaths and tumours among rotogravure printers exposed to toluene. British Journal of Industrial Medicine 47(6):372–379.


Tatham L, Tolbert P, Kjeldsberg C. 1997. Occupational risk factors for subgroups of non-Hodgkin’s lymphoma. Epidemiology 8(5):551–558.

Teschke K, Morgan MS, Checkoway H, Franklin G, Spinelli JJ, Van BG, Weiss NS. 1997. Surveillance of nasal and bladder cancer to locate sources of exposure to occupational carcinogens. Occupational and Environmental Medicine 54(6):443–451.

Teta MJ, Perlman GD, Ott MG. 1992. Mortality study of ethanol and isopropanol production workers at two facilities. Scandinavian Journal of Work, Environment and Health 18(2):90–96.

Thomas TL, Stewart PA, Stemhagen A, Correa P, Norman SA, Bleecker ML, Hoover RN. 1987. Risk of astrocytic brain tumors associated with occupational chemical exposures. A case-referent study. Scandinavian Journal of Work, Environment and Health 13(5):417–423.

Tomenson JA, Bonner SM, Heijne CG, Farrar DG, Cummings TF. 1997. Mortality of workers exposed to methylene chloride employed at a plant producing cellulose triacetate film base. Occupational and Environmental Medicine 54(7):470–476.

Tsai SP, Wen CP, Weiss NS, Wong O, McClellan WA, Gibson RL. 1983. Retrospective mortality and medical surveillance studies of workers in benzene areas of refineries. Journal of Occupational Medicine 25(9):685–692.


US Surgeon General. 1985 The Health Consequences of Smoking: Cancer and Chronic Lung Disease in the Workplace: A Report of the Surgeon General. Rockville, MD: Department of Health and Human Services.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Vamvakas S, Bruning T, Thomasson B, Lammert M, Baumuller A, Bolt HM, Dekant W, Birner G, Henschler D, Ulm K. 1998. Renal cell cancer correlated with occupational exposure to trichloroethene. Journal of Cancer Research and Clinical Oncology 124(7):374–382.

Van den Berghe H, Louwagie A, Broeckaert-Van Orshoven A, David G, Verwilghen R. 1979. Chromosome analysis in two unusual malignant blood disorders presumably induced by benzene. Blood 53(4):558–566.

van Steensel-Moll HA, Valkenburg HA, van Zanen GE. 1985. Childhood leukemia and parental occupation. A register-based case-control study. American Journal of Epidemiology 121(2):216–224.

Vaughan TL, Stewart PA, Davis S, Thomas DB. 1997. Work in dry-cleaning and the incidence of cancer of the oral cavity, larynx, and oesophagus. Occupational and Environmental Medicine 54(9):692–695.

Viadana E, Bross ID. 1972. Leukemia and occupations. Preventive Medicine 1(4):513–521.

Vineis P, Magnani C. 1985. Occupation and bladder cancer in males: A case-control study. International Journal of Cancer 35(5):599–606.


Walker JT, Bloom TF, Stern FB, Okun AH, Fingerhut MA, Halperin WE. 1993. Mortality of workers employed in shoe manufacturing. Scandinavian Journal of Work, Environment and Health 19(2):89–95.

Wallace J. 1996. Phenol. In: Kirk RE, Kroschwitz JI, Howe-Grant M, eds. Encyclopedia of Chemical Technology. Vol. 18. 4th ed. New York: John Wiley & Sons.

Waxweiler RJ, Smith AH, Falk H, Tyroler HA. 1981. Excess lung cancer risk in a synthetic chemicals plant. Environmental Health Perspectives 41:159–165.

Webler T, Brown HS. 1993. Exposure to tetrachloroethylene via contaminated drinking water pipes in Massachusetts: A predictive model. Archives of Environmental Health 48(5):293–297.

Weiderpass E, Pukkala E, Kauppinen T, Mutanen P, Paakkulainen H, Vasama-Neuvonen K, Boffetta P, Partanen T. 1999. Breast cancer and occupational exposures in women in Finland. American Journal of Industrial Medicine 36(1):48–53.

Wen CP, Tsai SP, Weiss NS, Gibson RL, Wong O, McClellan WA. 1985. Long-term mortality study of oil refinery workers. IV. Exposure to the lubricating-dewaxing process. Journal of the National Cancer Institute 74(1):11–18.

West RR, Stafford DA, Farrow A, Jacobs A. 1995. Occupational and environmental exposures and myelodysplasia: A case-control study. Leukemia Research 19(2):127–139.

West RR, Stafford DA, White AD, Bowen DT, Padua RA. 2000. Cytogenetic abnormalities in the myelodysplastic syndromes and occupational or environmental exposure. Blood 95(6):2093–2097.

Wiebelt H, Becker N. 1999. Mortality in a cohort of toluene exposed employees (rotogravure printing plant workers). Journal of Occupational and Environmental Medicine 41(12):1134–1139.

Wilcosky TC, Checkoway H, Marshall EG, Tyroler HA. 1984. Cancer mortality and solvent exposures in the rubber industry. American Industrial Hygiene Association Journal 45(12):809–811.

Wolf PH, Andjelkovich D, Smith A, Tyroler H. 1981. A case-control study of leukemia in the U.S. rubber industry. Journal of Occupational Medicine 23(2):103–108.

Wong O. 1987a. An industry wide mortality study of chemical workers occupationally exposed to benzene. I. General results. British Journal of Industrial Medicine 44(6):365–381.

Wong O. 1987b. An industry wide mortality study of chemical workers occupationally exposed to benzene. II. Dose response analyses. British Journal of Industrial Medicine 44(6):382–395.

Wong O. 1995. Risk of acute myeloid leukaemia and multiple myeloma in workers exposed to benzene. Occupational and Environmental Medicine 52(6):380–384.

Wong O, Harris F, Smith TJ. 1993. Health effects of gasoline exposure. II. Mortality patterns of distribution workers in the United States. Environmental Health Perspectives 101(Suppl 6):63–76.


Yin SN, Li GL, Tain FD, Fu ZI, Jin C, Chen YJ, Luo SJ, Ye PZ, Zhang JZ, Wang GC, et al. 1987. Leukaemia in benzene workers: A retrospective cohort study. British Journal of Industrial Medicine 44(2):124–128.

Yin SN, Li GL, Tain FD, Fu ZI, Jin C, Chen YJ, Luo SJ, Ye PZ, Zhang JZ, Wang GC, Zhang XC, Wu HN, Zhong QC. 1989. A retrospective cohort study of leukemia and other cancers in benzene workers. Environmental Health Perspectives 82:207–213.

Yin SN, Linet MS, Hayes RB, Li GL, Dosemeci M, Wang YZ, Chow WH, Jiang ZL, Wacholder S, Zhang WU, et al. 1994. Cohort study among workers exposed to benzene in China: I. General methods and resources. American Journal of Industrial Medicine 26(3):383–400.

Yin SN, Hayes RB, Linet MS, Li GL, Dosemeci M, Travis LB, Li CY, Zhang ZN, Li DG, Chow WH, Wacholder S, Wang YZ, Jiang ZL, Dai TR, Zhang WY, Chao XJ, Ye PZ, Kou QR, Zhang XC, Lin XF, Meng JF, Ding CY, Zho JS, Blot WJ. 1996a. A cohort study of cancer among benzene-exposed workers in China: Overall results. American Journal of Industrial Medicine 29(3):227–235.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×

Yin SN, Hayes RB, Linet MS, Li GL, Dosemeci M, Travis LB, Zhang ZN, Li DG, Chow WH, Wacholder S, Blot WJ. 1996b. An expanded cohort study of cancer among benzene-exposed workers in China. Benzene Study Group. Environmental Health Perspectives 104(Suppl 6):1339–1341.


Zack M, Cannon S, Loyd D, Heath CW Jr, Falletta JM, Jones B, Housworth J, Crowley S. 1980. Cancer in children of parents exposed to hydrocarbon-related industries and occupations. American Journal of Epidemiology 111(3):329–336.

Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 156
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 157
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 158
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 159
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 160
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 161
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 162
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 163
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 164
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 165
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 166
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 167
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 168
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 169
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 170
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 171
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 172
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 173
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 174
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 175
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 176
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 177
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 178
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 179
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 180
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 181
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 182
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 183
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 184
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 185
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 186
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 187
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 188
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 189
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 190
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 191
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 192
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 193
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 194
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 195
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 196
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 197
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 198
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 199
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 200
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 201
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 202
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 203
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 204
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 205
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 206
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 207
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 208
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 209
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 210
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 211
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 212
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 213
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 214
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 215
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 216
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 217
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 218
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 219
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 220
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 221
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 222
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 223
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 224
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 225
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 226
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 227
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 228
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 229
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 230
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 231
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 232
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 233
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 234
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 235
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 236
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 237
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 238
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 239
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 240
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 241
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 242
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 243
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 244
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 245
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 246
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 247
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 248
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 249
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 250
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 251
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 252
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 253
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 254
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 255
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 256
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 257
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 258
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 259
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 260
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 261
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 262
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 263
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 264
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 265
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 266
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 267
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 268
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 269
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 270
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 271
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 272
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 273
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 274
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 275
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 276
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 277
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 278
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 279
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 280
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 281
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 282
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 283
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 284
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 285
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 286
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 287
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 288
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 289
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 290
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 291
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 292
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 293
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 294
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 295
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 296
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 297
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 298
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 299
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 300
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 301
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 302
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 303
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 304
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 305
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 306
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 307
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 308
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 309
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 310
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 311
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 312
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 313
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 314
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 315
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 316
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 317
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 318
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 319
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 320
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 321
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 322
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 323
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 324
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 325
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 326
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 327
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 328
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 329
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 330
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 331
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 332
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 333
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 334
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 335
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 336
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 337
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 338
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 339
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 340
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 341
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 342
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 343
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 344
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 345
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 346
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 347
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 348
Suggested Citation:"6. Cancer and Exposure to Solvents." Institute of Medicine. 2003. Gulf War and Health: Volume 2: Insecticides and Solvents. Washington, DC: The National Academies Press. doi: 10.17226/10628.
×
Page 349
Next: 7. Neurologic Effects »
Gulf War and Health: Volume 2: Insecticides and Solvents Get This Book
×
Buy Paperback | $150.00 Buy Ebook | $119.99
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

Gulf War and Health, Volume 2, is the second in a series of congressionally-mandated studies by the Institute of Medicine that provides a comprehensive assessment of the available scientific literature on potential health effects of exposure to certain biological, chemical, and environmental agents associated with the Gulf War. In this second study, the committee evaluated the published, peer-reviewed literature on exposure to insecticides and solvents thought to have been present during the 1990-1991 war.

Because little information exists on actual exposure levels – a critical factor when assessing health effects – the committee could not draw specific conclusions about the health problems of Gulf War veterans. However, the study found some evidence, although usually limited, to link specific long-term health outcomes with exposure to certain insecticides and solvents.

The next phase of the series will examine the literature on potential health effects associated with exposure to selected environmental pollutants and particulates, such as oil-well fires and jet fuels.

  1. ×

    Welcome to OpenBook!

    You're looking at OpenBook, NAP.edu's online reading room since 1999. Based on feedback from you, our users, we've made some improvements that make it easier than ever to read thousands of publications on our website.

    Do you want to take a quick tour of the OpenBook's features?

    No Thanks Take a Tour »
  2. ×

    Show this book's table of contents, where you can jump to any chapter by name.

    « Back Next »
  3. ×

    ...or use these buttons to go back to the previous chapter or skip to the next one.

    « Back Next »
  4. ×

    Jump up to the previous page or down to the next one. Also, you can type in a page number and press Enter to go directly to that page in the book.

    « Back Next »
  5. ×

    Switch between the Original Pages, where you can read the report as it appeared in print, and Text Pages for the web version, where you can highlight and search the text.

    « Back Next »
  6. ×

    To search the entire text of this book, type in your search term here and press Enter.

    « Back Next »
  7. ×

    Share a link to this book page on your preferred social network or via email.

    « Back Next »
  8. ×

    View our suggested citation for this chapter.

    « Back Next »
  9. ×

    Ready to take your reading offline? Click here to buy this book in print or download it as a free PDF, if available.

    « Back Next »
Stay Connected!