4,451
Views
69
CrossRef citations to date
0
Altmetric
Review Article

Health effects research and regulation of diesel exhaust: an historical overview focused on lung cancer risk

, , , , &
Pages 1-45 | Received 20 Feb 2012, Accepted 03 May 2012, Published online: 04 Jun 2012

Figures & data

Table 1.  Key regulatory actions affecting diesel engine exhaust in the United States.

Figure 1.  Evolution of US heavy-duty diesel engine on-road emissions standards, expressed as grams PM or NOx emitted per brake-horsepower-hour (g/bhp-hr). Note that in 2004 two alternative standards were implemented: either a combined NOx+NMHC limit of 2.4 g/bhp-hr, or a NOx limit of 2.5 g/bhp-hr and a NMHC limit of 0.5 g/bhp-hr. See for additional details and citations for the emissions standards.

Figure 1.  Evolution of US heavy-duty diesel engine on-road emissions standards, expressed as grams PM or NOx emitted per brake-horsepower-hour (g/bhp-hr). Note that in 2004 two alternative standards were implemented: either a combined NOx+NMHC limit of 2.4 g/bhp-hr, or a NOx limit of 2.5 g/bhp-hr and a NMHC limit of 0.5 g/bhp-hr. See Table 1 for additional details and citations for the emissions standards.

Figure 2.  Chemical compositions of PM in NTDE (data from Khalek et al., Citation2011; based on averaged data for four 2007-model-year heavy-duty diesel engines, including three equipped with a diesel oxidation catalyst (DOC) and a catalyzed diesel particulate filter (c-DPF), and engine equipped with an exhaust diesel fuel burner and c-DPF) versus TDE (data from US EPA, Citation2002; for 1990s-era diesel engine technology) from heavy-duty diesel engines. PM mass emissions bars for NTDE and TDE derived from data compiled in Hesterberg et al. (2008) for diesel school buses with and without catalyzed DPFs (used in conjunction with ULSD), respectively. Note that there can be variability in PM emissions for diesel engine technologies considered to emit NTDE and TDE, such that data from other studies may differ from those in the figure. In general, as illustrated in these comparisons, not only is less PM emitted in NTDE on a per mile basis, but the emitted PM differs in composition from the PM emitted in TDE.

Figure 2.  Chemical compositions of PM in NTDE (data from Khalek et al., Citation2011; based on averaged data for four 2007-model-year heavy-duty diesel engines, including three equipped with a diesel oxidation catalyst (DOC) and a catalyzed diesel particulate filter (c-DPF), and engine equipped with an exhaust diesel fuel burner and c-DPF) versus TDE (data from US EPA, Citation2002; for 1990s-era diesel engine technology) from heavy-duty diesel engines. PM mass emissions bars for NTDE and TDE derived from data compiled in Hesterberg et al. (2008) for diesel school buses with and without catalyzed DPFs (used in conjunction with ULSD), respectively. Note that there can be variability in PM emissions for diesel engine technologies considered to emit NTDE and TDE, such that data from other studies may differ from those in the figure. In general, as illustrated in these comparisons, not only is less PM emitted in NTDE on a per mile basis, but the emitted PM differs in composition from the PM emitted in TDE.

Figure 3.  Average % reductions for DEP chemical classes relative to 2004 diesel technology engines for ACES testing of four post-2006 technology diesel engines (data from Khalek et al., Citation2011). ACES testing for 12 repeats of 16-h transient cycle developed at West Virginia University that covers a complete engine operation with active regeneration events. *Reductions in dioxins/furans are for comparison with 1998 technology engines.

Figure 3.  Average % reductions for DEP chemical classes relative to 2004 diesel technology engines for ACES testing of four post-2006 technology diesel engines (data from Khalek et al., Citation2011). ACES testing for 12 repeats of 16-h transient cycle developed at West Virginia University that covers a complete engine operation with active regeneration events. *Reductions in dioxins/furans are for comparison with 1998 technology engines.

Figure 4.  Average particle number emissions (note the logarithmic scale) for 2007 ACES engines (with and without c-DPF regeneration) versus a 2004 technology engine. As discussed in Khalek et al. (Citation2011), data for the 2007 ACES engines were based on 12 repeats of the 20-min federal test procedure transient cycle (FTP) or 12 repeats of the 16-h cycle, each for all four ACES engines and for sampling from an unoccupied animal exposure chamber set up on a constant volume sampler (CVS). Data for the 2004 technology engine were based on six repeats of the FTP transient cycle from a full flow CVS. All data are reported on a brake-specific emissions basis, which is defined by Khalek et al. (Citation2011) as the total emissions during a test interval over the work expressed in brake horsepower-hour.

Figure 4.  Average particle number emissions (note the logarithmic scale) for 2007 ACES engines (with and without c-DPF regeneration) versus a 2004 technology engine. As discussed in Khalek et al. (Citation2011), data for the 2007 ACES engines were based on 12 repeats of the 20-min federal test procedure transient cycle (FTP) or 12 repeats of the 16-h cycle, each for all four ACES engines and for sampling from an unoccupied animal exposure chamber set up on a constant volume sampler (CVS). Data for the 2004 technology engine were based on six repeats of the FTP transient cycle from a full flow CVS. All data are reported on a brake-specific emissions basis, which is defined by Khalek et al. (Citation2011) as the total emissions during a test interval over the work expressed in brake horsepower-hour.

Table 2.  Overview of reported exposure levels for DE-exposed worker groups and the general population based on EC measurements and predicted DEP concentrations.

Figure 5.  Median predicted shift-level elemental carbon (EC) concentrations for trucking industry workers by decade (1971–1980, 1981–1990, 1991–2000), as reported in Davis et al. (Citation2011). Job-specific concentrations are summarized, with multiple predictions for dockworkers corresponding to use of diesel-powered, propane-powered, and gasoline-powered forklifts and separate predictions for both mechanics and pickup & delivery drivers in warm versus cold climates. As discussed in Davis et al. (Citation2011), their modeling analysis provides evidence of substantial reductions in truckers’ DE exposures over the last three decades. LH stands for long-haul, while P&D stands for pickup-and-delivery.

Figure 5.  Median predicted shift-level elemental carbon (EC) concentrations for trucking industry workers by decade (1971–1980, 1981–1990, 1991–2000), as reported in Davis et al. (Citation2011). Job-specific concentrations are summarized, with multiple predictions for dockworkers corresponding to use of diesel-powered, propane-powered, and gasoline-powered forklifts and separate predictions for both mechanics and pickup & delivery drivers in warm versus cold climates. As discussed in Davis et al. (Citation2011), their modeling analysis provides evidence of substantial reductions in truckers’ DE exposures over the last three decades. LH stands for long-haul, while P&D stands for pickup-and-delivery.

Figure 6.  Histogram of predicted annual county-average ambient diesel particulate matter (DPM) concentrations for the US EPA National-Scale Air Toxics Assessment (NATA) modeling analyses of 1996 and 2005 year air pollutant emissions (data from US EPA, 2011). DPM emissions include both on-road and non-road emissions sources. County numbers (out of 3191 counties for the 1996 emission year modeling and 3221 counties for the 2005 emission year modeling; both including municipalities in Puerto Rico and counties in the US Virgin Islands) are provided above each bar. These data suggest a decline in ambient DE exposure levels between 1996 and 2005, although there have also been improvements in NATA methods (e.g. inventory improvements, modeling changes, background calculation revisions) over time that may affect the interpretation of any differences between the two NATA analyses.

Figure 6.  Histogram of predicted annual county-average ambient diesel particulate matter (DPM) concentrations for the US EPA National-Scale Air Toxics Assessment (NATA) modeling analyses of 1996 and 2005 year air pollutant emissions (data from US EPA, 2011). DPM emissions include both on-road and non-road emissions sources. County numbers (out of 3191 counties for the 1996 emission year modeling and 3221 counties for the 2005 emission year modeling; both including municipalities in Puerto Rico and counties in the US Virgin Islands) are provided above each bar. These data suggest a decline in ambient DE exposure levels between 1996 and 2005, although there have also been improvements in NATA methods (e.g. inventory improvements, modeling changes, background calculation revisions) over time that may affect the interpretation of any differences between the two NATA analyses.

Table 3.  Timeline of key DE health effects research milestones.

Table 4.  Summary of recent (post-US EPA Diesel HAD, 2001 and later) epidemiological studies of DE-lung cancer risk.

Figure 7.  Chart shows study-specific and overall pooled-study lung-cancer odds ratios (OR) and 95% confidence intervals (CIs) for the highest quartile of cumulative diesel exhaust exposure compared with never-exposed, adjusted for age, sex, cigarette pack-years, time-since-quitting smoking, and ever-employment in a “List A” job (from Olsson et al., Citation2011a). Studies are identified by locations, with study acronyms provided in parentheses. As summarized in our ,Olsson et al. (Citation2011a) pooled information from 11 European and Canadian case–control studies covering 13,304 cases, with exposures typically between the 1920s/1930s and the 1990s/2000s. As noted inOlsson et al. (Citation2011a), the symbol size reflects weighting from the random effects analysis. For global testing of the heterogeneity between the study ORs,Olsson et al. (Citation2011a) reported an overall I-squared (I2) of 13.8% (p = 0.292) and concluded that there was no significant heterogeneity.

Figure 7.  Chart shows study-specific and overall pooled-study lung-cancer odds ratios (OR) and 95% confidence intervals (CIs) for the highest quartile of cumulative diesel exhaust exposure compared with never-exposed, adjusted for age, sex, cigarette pack-years, time-since-quitting smoking, and ever-employment in a “List A” job (from Olsson et al., Citation2011a). Studies are identified by locations, with study acronyms provided in parentheses. As summarized in our Table 4,Olsson et al. (Citation2011a) pooled information from 11 European and Canadian case–control studies covering 13,304 cases, with exposures typically between the 1920s/1930s and the 1990s/2000s. As noted inOlsson et al. (Citation2011a), the symbol size reflects weighting from the random effects analysis. For global testing of the heterogeneity between the study ORs,Olsson et al. (Citation2011a) reported an overall I-squared (I2) of 13.8% (p = 0.292) and concluded that there was no significant heterogeneity.

Figure 8.  Possible mechanistic pathways leading to lung tumors in rats exposed by inhalation to protracted, high concentrations of poorly-soluble particles (adapted from Hesterberg et al., Citation2005 and HEI, Citation1995).

Figure 8.  Possible mechanistic pathways leading to lung tumors in rats exposed by inhalation to protracted, high concentrations of poorly-soluble particles (adapted from Hesterberg et al., Citation2005 and HEI, Citation1995).

Figure 9.  Impaired lung clearance in rats of 134Cs-radiolabeled particles inhaled after the end of 24-months DE exposure (for high, medium, and low DE exposure concentrations of 7.0, 3.5, and 0.35 μg/m3, respectively) and for a control population (0 mg/m3 DE exposure). Data points are means ± standard errors (SEs). From Wolff et al. (Citation1987).

Figure 9.  Impaired lung clearance in rats of 134Cs-radiolabeled particles inhaled after the end of 24-months DE exposure (for high, medium, and low DE exposure concentrations of 7.0, 3.5, and 0.35 μg/m3, respectively) and for a control population (0 mg/m3 DE exposure). Data points are means ± standard errors (SEs). From Wolff et al. (Citation1987).

Figure 10.  Relationship of normalized weekly exposure of rats to DEP versus rat lung tumor response (adapted from Mauderly and Garshick, Citation2009). Data from nine published studies with groups of 50 or more rats exposed ≥24 months to DE; data from the single chronic rat study published since the 1988 IARC DE review – Stinn et al. (Citation2005) – are specifically labeled. Lung tumor increases are shown (exposed minus controls). Dashed line represents control incidence (no net increase). Open circles represent exposed groups with no statistically significant increase above the control incidence. Closed circles represent exposed groups with a statistically significant increase above individual control group lung tumor incidence. In addition to the DEP study data, we have also plotted data for carbon black (CB) from Nikula et al. (Citation1995). Although Heinrich et al. (Citation1995) also included a CB exposure group and observed a 27% excess in lung tumor incidence (exposed minus controls), we did not include this data point in the figure since the weekly exposure rate of 990 mg-h/m3 is well outside the range of DEP exposure rates and would have thus distorted the figure scale.

Figure 10.  Relationship of normalized weekly exposure of rats to DEP versus rat lung tumor response (adapted from Mauderly and Garshick, Citation2009). Data from nine published studies with groups of 50 or more rats exposed ≥24 months to DE; data from the single chronic rat study published since the 1988 IARC DE review – Stinn et al. (Citation2005) – are specifically labeled. Lung tumor increases are shown (exposed minus controls). Dashed line represents control incidence (no net increase). Open circles represent exposed groups with no statistically significant increase above the control incidence. Closed circles represent exposed groups with a statistically significant increase above individual control group lung tumor incidence. In addition to the DEP study data, we have also plotted data for carbon black (CB) from Nikula et al. (Citation1995). Although Heinrich et al. (Citation1995) also included a CB exposure group and observed a 27% excess in lung tumor incidence (exposed minus controls), we did not include this data point in the figure since the weekly exposure rate of 990 mg-h/m3 is well outside the range of DEP exposure rates and would have thus distorted the figure scale.

Table 5.  Summary of pulmonary effects in different species related to high particle load (from Oberdörster, Citation1995).

Figure 11.  Summary of McDonald et al. (Citation2004b) findings on the relative toxicity in mice of acute inhalation exposures (6 h per day over 7 days) for a baseline uncontrolled, TDE emissions case (approximately 200 μg/m3 DEP) versus an emissions reduction case (low-sulfur fuel, catalyzed ceramic trap, 7 μg/m3). Expressed as relative responses to filtered air, findings are shown for four indicators of acute lung toxicity, namely respiratory syncytial virus (RSV) resistance, histopathology, lung inflammation (specifically, measurements of tumor necrosis factor-α (TNF-α)), and oxidative stress.

Figure 11.  Summary of McDonald et al. (Citation2004b) findings on the relative toxicity in mice of acute inhalation exposures (6 h per day over 7 days) for a baseline uncontrolled, TDE emissions case (approximately 200 μg/m3 DEP) versus an emissions reduction case (low-sulfur fuel, catalyzed ceramic trap, 7 μg/m3). Expressed as relative responses to filtered air, findings are shown for four indicators of acute lung toxicity, namely respiratory syncytial virus (RSV) resistance, histopathology, lung inflammation (specifically, measurements of tumor necrosis factor-α (TNF-α)), and oxidative stress.

Figure 12.  Comparison of total PM emissions (on a mass per-distance-traveled basis) and PM composition for light-duty automobile engine exhausts representative of TDE, NTDE, and GEE. All data based on particle composition measurements from Cheung et al. (Citation2009), who conducted emissions testing on a chassis dynamometer for light-duty vehicles operated using different aftertreatment configurations and a cold-start New European Driving Cycle (NEDC) and a series of Artemis cycles. Specific vehicle configurations include a Euro 4+ Honda Accord (2.2 L, i-CDTi) equipped with a ceramic-catalyzed diesel particulate filter (c-DPF), a closed-coupled oxidation catalyst (pre-cat), and exhaust gas recirculation (EGR), operated using low sulfur (<10 ppm) diesel fuel and lube oil with a sulfur content of 8900 ppm wt (considered to be NTDE); a Euro 3 Toyota Corolla (1.8 L) equipped with a three-way catalytic converter and operated using unleaded gasoline with a research octane number (RON) of 95 and fully synthetic lube oil (considered to be GEE); and a Euro 1 compliant Volkswagen Golf (TDI, 1.9 L) operated using diesel fuel with a nominal sulfur content of 50 ppm (considered to be TDE).

Figure 12.  Comparison of total PM emissions (on a mass per-distance-traveled basis) and PM composition for light-duty automobile engine exhausts representative of TDE, NTDE, and GEE. All data based on particle composition measurements from Cheung et al. (Citation2009), who conducted emissions testing on a chassis dynamometer for light-duty vehicles operated using different aftertreatment configurations and a cold-start New European Driving Cycle (NEDC) and a series of Artemis cycles. Specific vehicle configurations include a Euro 4+ Honda Accord (2.2 L, i-CDTi) equipped with a ceramic-catalyzed diesel particulate filter (c-DPF), a closed-coupled oxidation catalyst (pre-cat), and exhaust gas recirculation (EGR), operated using low sulfur (<10 ppm) diesel fuel and lube oil with a sulfur content of 8900 ppm wt (considered to be NTDE); a Euro 3 Toyota Corolla (1.8 L) equipped with a three-way catalytic converter and operated using unleaded gasoline with a research octane number (RON) of 95 and fully synthetic lube oil (considered to be GEE); and a Euro 1 compliant Volkswagen Golf (TDI, 1.9 L) operated using diesel fuel with a nominal sulfur content of 50 ppm (considered to be TDE).

Table 6.  Summary of DE/DEP hazard assessments conducted by regulatory agencies and authoritative bodies.

Supplemental material

Supplementary Material

Download PDF (511.2 KB)