Skip to main content
Environmental Health Perspectives logoLink to Environmental Health Perspectives
. 2005 Mar 16;113(8):934–946. doi: 10.1289/ehp.7938

Potential Role of Ultrafine Particles in Associations between Airborne Particle Mass and Cardiovascular Health

Ralph J Delfino 1, Constantinos Sioutas 2, Shaista Malik 3
PMCID: PMC1280331  PMID: 16079061

Abstract

Numerous epidemiologic time-series studies have shown generally consistent associations of cardiovascular hospital admissions and mortality with outdoor air pollution, particularly mass concentrations of particulate matter (PM) ≤2.5 or ≤10 μm in diameter (PM2.5, PM10). Panel studies with repeated measures have supported the time-series results showing associations between PM and risk of cardiac ischemia and arrhythmias, increased blood pressure, decreased heart rate variability, and increased circulating markers of inflammation and thrombosis. The causal components driving the PM associations remain to be identified. Epidemiologic data using pollutant gases and particle characteristics such as particle number concentration and elemental carbon have provided indirect evidence that products of fossil fuel combustion are important. Ultrafine particles < 0.1 μm (UFPs) dominate particle number concentrations and surface area and are therefore capable of carrying large concentrations of adsorbed or condensed toxic air pollutants. It is likely that redox-active components in UFPs from fossil fuel combustion reach cardiovascular target sites. High UFP exposures may lead to systemic inflammation through oxidative stress responses to reactive oxygen species and thereby promote the progression of atherosclerosis and precipitate acute cardiovascular responses ranging from increased blood pressure to myocardial infarction. The next steps in epidemiologic research are to identify more clearly the putative PM casual components and size fractions linked to their sources. To advance this, we discuss in a companion article (Sioutas C, Delfino RJ, Singh M. 2005. Environ Health Perspect 113:947–955) the need for and methods of UFP exposure assessment.

Keywords: cardiovascular diseases, cytokines, diesel, epidemiology, oxidative stress, particle size, toxic air pollutants


Coronary heart disease (CHD) is the leading cause of death and hospitalization among adults 65 or more years of age (Desai et al. 1999), which makes the identification of preventable causes for heart disease morbidity and mortality an important research goal. Numerous epidemiologic time-series studies have shown generally consistent associations of outdoor (ambient) air pollution with cardiovascular hospital admissions (Burnett et al. 1995, 1997a, 1997b, 1999; D’Ippoliti et al. 2003; Le Tertre et al. 2002; Linn et al. 2000; Mann et al. 2002; Morris et al. 1995; Peters et al. 2001a; Poloniecki et al. 1997; Samet et al. 2000a; Schwartz 1999; Schwartz and Morris 1995; Zanobetti and Schwartz 2001; Zanobetti et al. 2000a, 2000b). Consistent associations of ambient air pollution have also been found with cardiovascular mortality (Clancy et al. 2002; Dockery et al. 1993; Goldberg et al. 2001a, 2001b; Hoek et al. 2001; Kwon et al. 2001; Laden et al. 2000; Pope et al. 2004a; Rossi et al. 1999; Samet et al. 2000b; Schwartz et al. 1996; Wichmann et al. 2000; Zanobetti et al. 2003). The National Research Council (NRC) Committee on Research Priorities for Airborne Particulate Matter has identified research needed to explain the morbidity and mortality associations in the time-series studies (NRC 1998, 1999, 2001, 2004). One priority is to identify the pathophysiologic mechanisms and causal pollutant components driving these associations (Seaton et al. 1995).

The causal components driving the relationship between particulate matter (PM) and cardiovascular morbidity and mortality remain to be identified. Historically, the difficulty in accomplishing this in epidemiologic studies is related to the common use of ambient air pollution data from monitoring stations located at central regional sites. This has led to both exposure misclassification and high correlations between different pollutants. Both of these problems can be addressed with measurements of personal and/or microenvironmental exposures (Sarnat et al. 2000, 2001). Another problem is that the importance of particle size and chemistry has been limited by reliance on the same government monitoring data. In the United States, these data generally include only particle mass concentrations in air at two particle size cuts, PM10 (PM ≤ 10 μm in aerodynamic diameter) and more recently PM2.5 (PM ≤ 2.5 μm). However, there is sufficient reason to believe that ultrafine particles (UFPs; PM < 0.1 μm) are important in morbidity and mortality associations otherwise attributed to larger-size fractions.

Major characteristics of UFPs that support their potential importance include a high pulmonary deposition efficiency, magnitudes higher particle number concentration than larger particles, and thus a much higher surface area. The UFP’s surface can carry large amounts of adsorbed or condensed toxic air pollutants (oxidant gases, organic compounds, transition metals) (Oberdörster 2001). Many of these toxic air pollutants have been identified as having pro-inflammatory effects in part through the action of reactive oxygen species (ROS), but relevant exposure data are rarely available to epidemiologists. Available surrogate measures of fossil fuel combustion such as elemental carbon (EC) or black smoke are of some use in this regard. Results from a study in southern California showed that a large proportion of urban UFPs is made up of primary combustion products from mobile source emissions (particularly diesel and automobile exhaust) and includes organic compounds, EC, and metals (Kim et al. 2002). Because exposure to mobile emissions can be variable across short distances and depends on personal activity patterns, assessing such exposures requires methods that go beyond the use of government monitoring data alone. These issues regarding the characteristics of UFPs are more thoroughly discussed in a companion article (Sioutas et al. 2005).

In the present review we discuss evidence for adverse effects of air pollution on cardiovascular health with an emphasis on findings that suggest a role for UFPs and related toxic air pollutant components. To date, there are fewer direct epidemiologic data on UFPs. We review studies using other particle size fractions, other particle measurements such as black smoke, and gas-phase pollutants to provide a rationale for investigations of UFPs. The focus of this article is on epidemiologic studies that have followed individual subjects over time. Several excellent reviews of experimental data and methods can be found elsewhere (Donaldson et al. 2001; Utell et al. 2002).

Evidence of Causal Pollutant Components in Epidemiologic Time-Series, Cohort, and Cross-Sectional Studies

The National Morbidity, Mortality, and Air Pollution Study (NMMAPS) is the largest of the air pollution time-series studies to date (Samet et al. 2000a, 2000b). Results show positive associations of PM10 with cardio-pulmonary mortality and with hospital admissions for cardiovascular disease, chronic obstructive pulmonary disease (COPD), and pneumonia in patients 65 or more years of age living in varied environments across up to 90 cities in the United States. A subsequent analysis to correct for statistical errors showed an increase of 0.34% [95% confidence interval (95% CI), 0.1–0.57] in combined cardio-respiratory mortality for each 10 μg/m3 of air increase in PM10 (Dominici et al. 2003). Another reanalysis of hospitalizations in 14 U.S. cities by Janssen et al. (2002) broke down the PM10 concentrations using information on source categories. The authors found that for cardiovascular admissions, and to a lesser extent COPD admissions, PM10 from highway vehicle and diesel emissions and from oil combustion showed the strongest associations with the most stable regression coefficients in co-regressions with other source categories. These findings are supported by an analysis of PM data collected for the Harvard Six Cities Study (Dockery et al. 1993) by Laden et al. (2000) using elemental profiles of PM2.5 samples. They showed that associations between daily total mortality and mobile source (largely traffic related) particles for the six metropolitan areas were twice those for sulfate-rich coal combustion particles. This difference was most clearly demonstrated for deaths from CHD.

Additional information regarding causal pollutant components has come from analyses of ambient gaseous air pollutants under U.S. federal regulation [carbon monoxide, nitrogen dioxide, sulfur dioxide, and ozone]. These pollutants can be strongly correlated with PM in ambient air. A European study by Katsouyanni et al. (2001) of 29 cities showed a positive association between total mortality and PM10 and that this association was not confounded by SO2 or O3. However, they did find that in cities with higher versus lower average NO2, the association with PM10 was significantly greater (0.80% vs. 0.19% increase in mortality per 10 μg/m3 PM10, respectively). The NMMAPS study found that PM10 associations with mortality were largely independent of NO2, SO2, and O3 (Samet et al. 2000a). Goldberg et al. (2001a, 2001b), Moolgavkar (2000), and Venners et al. (2003) have also found robust associations between cardiovascular mortality and pollutant gases that often were stronger than particle associations. In a time-series study of the Los Angeles air basin, Linn et al. (2000) found that significant associations of daily cardiovascular hospital admissions were strongest for CO, followed by NO2, and then much weaker associations for PM10, but daily PM data were limited by fewer stations. Morris et al. (1995) and Morris and Naumova (1998) found that hospital admissions for congestive heart failure (CHF) were associated with CO independent of other gaseous pollutants in several large U.S. cities. Mann et al. (2002) also found significant associations of daily CHD hospital admissions with NO2 and CO in Los Angeles, particularly among cases with a secondary diagnosis of CHF or arrhythmia. Lin et al. (2003) found that an interquartile range increase in CO was associated with an increase of 6.4% in daily angina and acute myocardial infarction (MI) emergency department visits in São Paulo, Brazil. A time-series study of seven European areas found cardiovascular hospital admissions, especially CHD, were associated with SO2 (Sunyer et al. 2003). Associations between gases and hospital admissions for CHD and CHF have been found in several other studies (e.g., Burnett et al. 1997b, 1999; Koken et al. 2003; Morris et al. 1995, 1998).

Some of the time-series investigators have hypothesized that pollutant gases could be acting as indicators for a causal mixture of pollutants, including PM-related components. Ambient CO is highly correlated with UFPs near combustion sources such as freeways (discussed more fully below). Although it is possible that some of the effects detected with CO are due to the formation of carboxyhemoglobin in the blood and carboxymyoglobin in muscle, reported ambient concentrations are low (< 6 ppm). A postulated mechanism for increased susceptibility to low CO doses is the attainment of a nominal threshold of reduced O2 transport to the heart and further compromised cardiac myoglobin, particularly in CHF patients (McGrath 2000).

Additional evidence of causal components linked to UFPs comes from European studies that have used a nongravimetric PM measure called black smoke, which is roughly representative of EC. Le Tertre et al. (2002) conducted a time-series analysis of cardiovascular hospital admissions in eight European cities and found that CHD admissions were associated with PM10 and black smoke. The association with PM10, but not with black smoke, was reduced by adding CO to the model and eliminated by adding NO2. Both Le Tertre et al. (2002) and the European study by Katsouyanni et al. (2001) reported above hypothesized that their results were attributable to traffic exhaust and its consequent high emissions of CO, NO2, black smoke, and air toxics. It is relevant to point out that traffic exhaust, particularly from diesel engines, is a major contributor to UFP mass in urban areas (Kittelson 1998; Tobias et al. 2001), and in general, UFPs are both strongly linked to mobile source emissions and laden with toxic constituents (Kim et al. 2002; Shi et al. 2001).

Although time-series investigations have provided important information regarding the overall public health impact of ambient air pollutants on severe outcomes such as mortality, studies of individual subjects have provided insights into the underlying acute or chronic exposure–response relationships. Below we review studies of individuals using various epidemiologic designs, including cohort and panel studies, focusing only on findings for cardiovascular outcomes. Details for selected studies are presented in Table 1 and follow the discussion in the text.

Table 1.

Cardiovascular effectsa associated with personal and ambient air pollution exposure: selected studies.

Studies Design and population Outcomes Findings for PM mass and components Findings for gases
Cohort and cross-sectional studies
Dockery et al. 1993 Cohort study examining ambient air pollution exposure and mortality in 8,111 adults in six U.S. cities with 14–16 years of follow-up Cardiopulmonary mortality Compared with the least polluted city, the most polluted city had an adjusted RR for cardiopulmonary mortality of 1.37 (95% CI, 1.11–1.68) No association with O3, but SO2 and NO2 tracked between-city trends in PM concentrations
Pope et al. 2004a Cohort study examining ambient PM exposure and cardiovascular mortality in 319,000–500,000 persons in the Cancer Prevention Study II, with 16 years of follow-up across U.S. urban areas Cardiovascular mortality: ischemic heart disease, dysrhythmias, heart failure, and cardiac arrest A 10-μg/m3 increase in PM2.5 was associated with 8–18% increases in mortality due to ischemic heart disease, dysrhythmias, heart failure, and cardiac arrest Not assessed
Abbey et al. 1999 Cohort study examining ambient PM10 exposure, total suspended sulfates, SO2, O3, and NO2 in relation to mortality in 6,338 non-smoking California Seventh-Day Adventists with 19 years of follow-up Cardiopulmonary mortality No associations No associations
Hoek et al. 2002 Cohort study examining ambient traffic-related air pollutant exposure (black smoke, NO2) and cause-specific mortality in 5,000 persons with 8 years of follow-up in the Netherlands Cohort Study on Diet and Cancer Cardiopulmonary mortality Cardiopulmonary mortality was associated with living near high traffic density (100 m to freeway or 50 m to major urban road) adjusted RR = 1.95 (95% CI, 1.09–3.52) and was associated with an increase of 10 μg/m3 black smoke from background (central sites) plus local sources (street proximity), RR = 1.71 (95% CI, 1.10–2.67) Cardiopulmonary mortality was associated with an increase of 30 μg/m3 background plus local NO2, RR 1.81 (95% CI, 0.98–3.34)
Künzli et al. 2004 Cross-sectional study on the relationship between ambient PM2.5 and CIMT, using baseline data from two clinical trials in Los Angeles; annual mean PM2.5 exposure was estimated using data from 23 monitoring stations linked to home addresses with geostatistical models CIMT For each increase of annual mean 10 μg/m3 PM2.5, CIMT increased by 5.9% (95% CI, 1–11%); adjustment for age reduced the coefficients, but further adjustment for covariates indicated robust estimates in the range of 3.9–4.3% Estimates for O3 linked to ZIP code centroids were positive in relation to CIMT but not significant and smaller than PM2.5
Cardiac ischemia and related outcomes
Pekkanen et al. 2002 Panel study examining ambient PM, NO2, and CO exposure and ischemia during 342 submaximal exercise tests in 45 subjects with CHD in Helsinki, Finland ECG ST segment depression > 0.1 mV Increased risk for ST depression (72 events) was associated with a change of lag-2 1,000 particles/cm3 NC0.1–1, OR = 3.29 (95% CI, 1.57–6.92), 10 μg/m3 PM2.5, OR = 2.84 (95% CI, 1.42–5.66), and 10,000 UFP/cm3 NC0.01–0.1, OR = 3.14 (95% CI, 1.56–6.32); UFPs were independent of PM2.5 NO2 and CO were also associated with an increased risk for ST depression.
de Hartog et al. 2003 Panel study examining ambient exposure to PM and NO2, SO2, and CO in relation to HRV and BP in 131 subjects with CHD in Helsinki, Finland; Amsterdam, the Netherlands; and Erfurt, Germany Cardiorespiratory symptoms: chest pain, shortness of breath, avoidance of activities A 10-μg/m3 increase in PM2.5 associated with shortness of breath, OR = 1.12 (95% CI, 1.02–1.24) and avoidance of activities, OR = 1.10 (95% CI, 1.01–1.19) Not assessed
Peters et al. 2004 Case-crossover study examining ambient traffic-related air pollution exposure and MI in 691 subjects from the Augsburg Myocardial Infarction Registry who had survived 24 hr postinfarct; time–activity diary data on activities during the 4 days before symptom onset were used to assess traffic exposures MI Exposure to traffic was associated with onset of MI 1 hr afterward, OR = 2.92 (95% CI, 2.22–3.83); a significant association was also seen for exposure to traffic 2 hr before onset, and there was evidence for effects up to 6 hr; key exposures influencing overall associations with traffic included times spent in cars and in public transportation; associations changed minimally, adjusting for exercise, and there was no confounding by reports of extreme anger or joy As with PM, gases were not directly assessed, but traffic exposures involve pollutant gases as well as particles
Blood pressure (BP)
Linn et al. 1999 Panel study in Los Angeles, California, examining BP and lung function in 30 subjects with COPD, with only 4 consecutive days of air sampling: personal exposure to PM2.5, indoor and outdoor home PM2.5 and PM10, and ambient PM10, O3, NO2, and CO BP Systolic BP increased 0.172 mm Hg for every 1-μg/m3 increase in ambient lag-1 PM10 (p = 0.006); diastolic BP increased 0.095 mm Hg for every 1-μg/m3 increase in PM10 (p = 0.03); outdoor home PM10 was similarly associated with BP, but no significant associations were reported for PM2.5 or any indoor or personal PM measurement No association of BP with exposure to central site O3, NO2, or CO
Brauer et al. 2001 Panel study examining personal exposure over 7 non-consecutive days to PM2.5 and sulfate, and ambient exposure to PM2.5, PM10 , sulfate, and gaseous pollutants, in relation to BP, HRV, and lung function in 16 COPD patients in Vancouver, Canada BP, HRV, SVE Weak associations were observed between particle concentrations and increased SVE and with decreased systolic BP; ambient PM10 had the largest effect on cardiovascular end points and the only statistically significant association (SVE); use of personal exposure measurements did not show a larger or more consistent effect CO was inversely associated with systolic BP and reduced estimates for ambient PM
Ibald-Mulli et al. 2001 Retrospective analysis examining the relationship between ambient air pollution exposure (TSP, SO2, and CO) and BP in 2,607 men and women 25–64 years of age from a general population survey in Augsburg, Germany Systolic BP A 90-μg/m3 increase in TSP was associated with an increase in systolic BP of 1.79 mm Hg (95% CI, 0.63–2.95); in subgroups with high plasma viscosity levels or increased HR, systolic BP increased by 6.93 mm Hg (95% CI, 4.31–9.75) and 7.76 mm Hg (95% CI, 5.70–9.82) in association with TSP, respectively An 80-μg/m3 increase in SO2 was associated with an increase in systolic BP of 0.74 mm Hg (95% CI, 0.08–1.40)
Ibald-Mulli et al. 2004 Panel study examining ambient exposure to PM and NO2, SO2, and CO in relation to HRV and BP in 131 subjects with CHD in Helsinki, Finland; Amsterdam, the Netherlands and Erfurt, Germany BP and HR A small decrease in systolic BP (−0.72 mm Hg; 95% CI, −1.92 to 0.49) and diastolic BP (−0.70 mm Hg; 95% CI, −0.02 to −1.38) was found to be associated with a 5-day average increase of 10,000 UFPs/cm3 (NC0.01–0.1); slightly stronger and more significant associations were found for accumulation mode particle number concentration (NC0.1–1.0), but smaller associations were found for a 10 μg/m3 increase in PM2.5 mass; small decreases in HR were also found for PM exposures The magnitude and significance of inverse BP associations with CO were similar to those of PM0.1–1.0; a small decrease in HR (−0.40 beats/min; 95% CI, −0.82 to 0.01) was found for an increase of lag-1, 5 μg/m3 SO2
Zanobetti et al. 2004 Panel study examining ambient PM2.5, O3, NO2, SO2, and CO in relation to BP among 62 patients with preexisting heart disease using data from 631 repeated visits for cardiac rehabilitation in Boston BP Increasing from the 10th to the 90th percentile in 5-day mean PM2.5 (10.5 μg/m3) resulted in increases of 2.8 mm Hg (95% CI, 0.1–5.5) in systolic, 2.7 mm Hg (95% CI, 1.2–4.3) in diastolic, and 2.7 mm Hg (95% CI, 1.0–4.5) in mean arterial BP; black carbon was associated with diastolic BP Diastolic BO was associated with 120-hr average SO2 (3.9% increase; 95% CI, 0.3–76), O3 (2.7% increase; 95% CI, 0.02–5.4)
Autonomic control of cardiac rhythm
Holguin et al. 2003 Panel study in Mexico City examining indoor and outdoor nursing home measurements of PM2.5 and ambient exposure to O3, NO2, CO, and SO2 in relation to HRV in 34 elderly residents followed every other day for 3 months; personal PM2.5 was predicted using indoor and outdoor home PM2.5 plus time–activity data HRV, frequency Domain A 10-μg/m3 increase in predicted personal PM2.5 was associated with a 5.0% decrease in high-frequency HRV (β = −0.049; 95% CI, −0.090 to −0.007); associations with indoor PM2.5 were stronger than outdoor home PM2.5; among 13 subjects with hypertension, the association with predicted personal PM2.5 was stronger (−7.1%) O3 was inversely associated with high-and low-frequency HRV among 13 subjects with hypertension (2% decrease per 10 ppb O3), but this association was confounded by PM2.5
Pope et al. 2004b Panel study of ambient exposure to PM and HRV and blood markers in 88 elderly subjects living in Salt Lake City and Provo/Orem, Utah HRV A 100-μg/m3 increase in PM2.5 was associated with a 35 (SE = 8) msec decrease in SDNN and a 42 (SE = 11) msec decrease in r-MSSD Not assessed
Autonomic control of cardiac rhythm
Magari et al. 2001, 2002a, 2002b Panel study examining personal exposure to PM in relation to HRV in 20 (Magari et al. 2002a), 40 (Magari et al. 2001), and 39 (Magari et al. 2002b) healthy boilermakers exposed to welding fumes and residual oil fly ash HRV Each 100-μg/m3 increase in 3-hr average PM2.5 (laser photometer light scatter) was associated with a 1.4% (95% CI, −2.1 to −0.6%) decrease in 5-min SDNN in the 20 subjects (Magari et al. 2002a); in the 40 subjects, each 1-mg/m3 increase in 4-hr average PM2.5 was associated with a 2.66% (95% CI, −3.75 to −1.58%) decrease in 5-min SDNN SDNN (Magari et al. 2001); however, in 39 of these 40 subjects, PM2.5 metals on filters, lead and vanadium, were associated with an increase in workday average of the 5-min SDNN (Magari et al. 2002b) Not assessed
Riediker et al. 2004 Panel study of in-vehicle exposure to PM and HRV and blood markers of inflammation in 9 healthy male North Carolina Highway Patrol troopers HRV In-vehicle 10-μg/m3 PM2.5 increase was associated with increased ectopic beats throughout exposure (20%, p = 0.005); PM2.5 was positively associated with heart beat cycle length (6%, p = 0.01) as well as HF HRV and SDNN the next morning after exposure NO2 and CO were not significant
Chan et al. 2004 Panel study in Taipei, Taiwan, examining personal exposure to sub-micrometer particles and HRV over one 16-hr daytime period in 9 young healthy adults 19–29 years of age (2 females) and 10 older male subjects 42–97 years of age with lung function impairments (FEV1/FVC < 85%) HRV Personal exposure to NC0.02–1 was associated with decreased in both time-domain and frequency-domain HRV indices; in young subjects, a 10,000 particles/cm3 increase in the last 1–4 hr average NC0.02–1 was associated with 0.68–1.35% decrease in SDNN, 1.85–2.58% decrease in r-MSSD; in the older panel they found 10,000-particles/cm3 increase in the last 1- to 3-hr average NC0.02–1 was associated 1.72–3.00% decreases in SDNN and 2.72–4.65% decreases in r-MSSD; there were similar associations for high- and low-frequency domain indices Not assessed
Tarkiainen et al. 2003 Panel study in Kuopio, Finland, examining personal exposure to carbon monoxide and HRV in 6 subjects with CHD followed for three separate 24-hr ambulatory monitoring periods HRV Not assessed r-MSSD increased by 2.4 msec (p = 0.03) with exposure to CO (> 2.7 ppm)
Peters et al. 2000 Panel study of arrhythmias in 100 subjects in eastern Massachusetts with implanted defibrillators (63,628 person-days of follow-up) with ambient measurements of PM mass, black carbon, NO2, CO, O3, and SO2 Defibrillator discharge interventions for ventricular tachycardias or fibrillation (33 subjects with at least one) Only 6 subjects with ≥ 10 defibrillator discharges had increased arrhythmias associated with black carbon and PM2.5, which showed a weaker association; both PM metrics were confounded by NO2, but the effect estimate of NO2 was unchanged 26-ppb increase in NO2 lagged 1 day was associated with increased defibrillator interventions in the full panel (OR = 1.8; 95% CI, 1.1–2.9). Subjects with ≥ 10 defibrillator discharges had increased arrhythmias associated with CO and NO2 across several lags
Systemic inflammation and thrombosis
Seaton et al. 1999 Panel study examining 3-day personal exposure estimated (from a one 24-hr personal exposure measurement) and city center ambient exposure to PM10 in relation to hematologic factors in 112 elderly subjects in Belfast and Edinburgh, UK Hematologic factors: hemoglobin, packed red cells, red blood cell count, platelets, white blood cell count, CRP, fibrinogen, factor VII, IL-6 An increase of 100 μ/m3 in personal PM10 and ambient PM10 exposure resulted in significant decreased mean percentage changes of ≤ 1% in hemoglobin concentration, packed cell volume, and red blood cell count; only personal PM10 was associated with an 11% decrease in platelets and a 7% decrease in factor VII; CRP increased with ambient PM10 (+147%; 95% CI, 20–477), but not with personal PM (p = 0.73); fibrinogen decreased with ambient PM10 (−9%; 95% CI, −19 to 0) Not assessed
Systemic inflammation and thrombosis
Schwartz 2001 Cross-sectional study examining the relationship between ambient PM10, NO2, SO2, and blood biomarkers using data from a cohort study (NHANES III) Fibrinogen, and platelet and white blood cell counts For an interquartile range change in PM10 (26 μg/m3), the relative odds for being above the 90th percentile of fibrinogen was 1.77 (95% CI, 1.26–2.49); platelets, 1.27 ( 95% CI, 0.97–1.67); and white blood cells, 1.64 (95% CI, 1.17–2.30) SO2 was positively associated with white cell counts, and NO2 with platelet counts and fibrinogen, but both gases were confounded by PM10
Pekkanen et al. 2000 Cross-sectional study examining the association between ambient PM10, NO2, CO, SO2, O3, and fibrinogen among 7,205 subjects in London at baseline enrollment in a cohort study Fibrinogen No association between PM10 and fibrinogen was seen after adjustment for confounders NO2 increase from the 10th to the 90th percentile was associated with a 1.5% higher fibrinogen concentration (95% CI, 0.4–2.5%); similar increase for CO resulted in 1.5% higher fibrinogen concentration (95% CI, 0.5–2.5%); no association with SO2 or O3
Peters et al. 1997a, 2001b Cohort study in Augsburg, Germany, examining relationships of ambient TSP, SO2, and CO exposure to CRP in 631 men 45–64 years of age with no history of MI at their baseline assessment; two CRP measurements were 3 years apart CRP An increase of 26 μg/m3 (5-day mean) in TSP increased the odds of observing a CRP level above the 80th percentile, OR = 1.31 (95% CI, 1.09–1.56); CRP and plasma viscosity (Peters et al. 1997a) were increased during an air pollution episode in 1985 An increase of 30 μg/m3 (5-day mean) in SO2 increased the odds of observing a CRP level above the 90th percentile, OR = 1.24 (95% CI, 1.03–1.49)
Pope et al. 2004b Panel study of ambient exposure to PM and HRV and blood markers in 88 elderly subjects living in Salt Lake City and Provo/Orem, Utah CRP, white blood cell count, whole blood viscosity, granulocytes, lymphocytes, monocytes, basophils, eosinophils, red blood cells, platelets A 100-μg/m3 increase in PM2.5 was associated with a 0.81 (SE = 0.17) mg/dL increase in CRP; one subject’s data had a strong influence on estimates; there was no association with other outcomes Not assessed
Riediker et al. 2004 Panel study of in-vehicle exposure to PM and HRV and blood markers of inflammation in 9 healthy male North Carolina Highway Patrol troopers CRP, plasminogen, von Willebrand factor, lymphocyte count, lymphocytes, neutrophils, hematocrit, red blood cell indices, uric acid In-vehicle 10-μg/m3 PM2.5 increase was associated with decreased lymphocytes (−11%, p = 0.03), increased red blood cell indices (1%, p = 0.03), neutrophils (6%, p = 0.04), CRP (32%, p = 0.02), and von Willebrand factor (12%, p = 0.02) NO2 and CO were not significant
Systemic inflammation and thrombosis
Schwartz 2001 Cross-sectional study examining the relationship between ambient PM10, NO2, SO2, and blood biomarkers using data from a cohort study (NHANES III) Fibrinogen, and platelet and white blood cell counts For an interquartile range change in PM10 (26 μg/m3), the relative odds for being above the 90th percentile of fibrinogen was 1.77 (95% CI, 1.26–2.49); platelets, 1.27 ( 95% CI, 0.97–1.67); and white blood cells, 1.64 (95% CI, 1.17–2.30) SO2 was positively associated with white cell counts, and NO2 with platelet counts and fibrinogen, but both gases were confounded by PM10
Pekkanen et al. 2000 Cross-sectional study examining the association between ambient PM10, NO2, CO, SO2, O3, and fibrinogen among 7,205 subjects in London at baseline enrollment in a cohort study Fibrinogen No association between PM10 and fibrinogen was seen after adjustment for confounders NO2 increase from the 10th to the 90th percentile was associated with a 1.5% higher fibrinogen concentration (95% CI, 0.4–2.5%); similar increase for CO resulted in 1.5% higher fibrinogen concentration (95% CI, 0.5–2.5%); no association with SO2 or O3
Peters et al. 1997a, 2001b Cohort study in Augsburg, Germany, examining relationships of ambient TSP, SO2, and CO exposure to CRP in 631 men 45–64 years of age with no history of MI at their baseline assessment; two CRP measurements were 3 years apart CRP An increase of 26 μg/m3 (5-day mean) in TSP increased the odds of observing a CRP level above the 80th percentile, OR = 1.31 (95% CI, 1.09–1.56); CRP and plasma viscosity (Peters et al. 1997a) were increased during an air pollution episode in 1985 An increase of 30 μg/m3 (5-day mean) in SO2 increased the odds of observing a CRP level above the 90th percentile, OR = 1.24 (95% CI, 1.03–1.49)
Pope et al. 2004b Panel study of ambient exposure to PM and HRV and blood markers in 88 elderly subjects living in Salt Lake City and Provo/Orem, Utah CRP, white blood cell count, whole blood viscosity, granulocytes, lymphocytes, monocytes, basophils, eosinophils, red blood cells, platelets A 100-μg/m3 increase in PM2.5 was associated with a 0.81 (SE = 0.17) mg/dL increase in CRP; one subject’s data had a strong influence on estimates; there was no association with other outcomes Not assessed
Riediker et al. 2004 Panel study of in-vehicle exposure to PM and HRV and blood markers of inflammation in 9 healthy male North Carolina Highway Patrol troopers CRP, plasminogen, von Willebrand factor, lymphocyte count, lymphocytes, neutrophils, hematocrit, red blood cell indices, uric acid In-vehicle 10-μg/m3 PM2.5 increase was associated with decreased lymphocytes (−11%, p = 0.03), increased red blood cell indices (1%, p = 0.03), neutrophils (6%, p = 0.04), CRP (32%, p = 0.02), and von Willebrand factor (12%, p = 0.02) NO2 and CO were not significant

Abbreviations: FEV1/FVC, forced expiratory volume in 1 sec/forced vital capacity; HF, high frequency; RR, relative risk; SVE, supraventricular ectopic heartbeat.

a

The focus is on cardiovascular outcomes. Although some studies may have examined other outcomes, they are not reported.

Time-series studies have provided evidence for acute effects of air pollutants on cardiovascular morbidity and mortality. However, there are still gaps in the literature regarding chronic health impacts from long-term pollutant exposures. Cohort studies are best suited to address this gap. Dockery et al. (1993) reported evidence from the Harvard Six Cities Study that ambient PM2.5 was associated with risk of cardiopulmonary mortality in a cohort of 8,111 adults (Table 1). Pope et al. (2004a) used 16 years of data from more than 500,000 adults in 151 U.S. cities that participated in the Cancer Prevention Study II of the American Cancer Society. The authors found that a 10-μg/m3 elevation in PM2.5 was associated with 8–18% increases in mortality due to ischemic heart disease, dysrhythmias, heart failure, and cardiac arrest. Mortality from various respiratory causes was not associated with PM2.5 (Table 1). In contrast, a cohort study of 6,338 Seventh Day Adventists living in California found associations of long-term exposure to PM and O3 with respiratory mortality but not with cardiovascular mortality (Abbey et al. 1999) (Table 1). Differences in findings might be due to exposure misclassification from the use of central regional air pollutant data. Hoek et al. (2002) tried to address this issue by evaluating effects of traffic exposures near the home in a cohort study of 5,000 adults followed 8 years in the Netherlands (Table 1). They showed that living near a major road was more strongly associated with cardiopulmonary mortality than with ambient background air pollutant levels. This finding suggests that pollutants more closely associated with traffic, which include UFPs and associated toxic air pollutants, could be causal components in the mortality associations.

Künzli et al. (2004) conducted a cross-sectional study of 798 healthy adults with elevated low density lipoprotein (LDL) cholesterol or homocysteine living on Los Angeles (Table 1). Subjects were in a dietary supplement clinical trial with ultrasound data on carotid intima-media thickness (CIMT) as an estimate of atherosclerosis. Exposure included an estimate using geostatistical models to link subject address to annual mean PM2.5 from 23 local air-monitoring stations. They found positive associations between CIMT and PM2.5, adjusting for host risk factors. Associations were larger for women, older subjects (≥ 60), subjects on lipid-lowering medications, and never smokers.

Evidence for Pathophysiologic Mechanisms and Causal Components in PM-Related Cardiovascular Effects

The following section looks at epidemiologic panel studies designed to evaluate the relationship between repeated air pollutant exposures and cardiovascular outcomes in individual subjects. We augment this discussion with a few selected human clinical studies that extend the panel study findings using controlled exposures, particularly those that aim to replicate ambient air mixtures. The discussion is divided by related groups of cardiovascular outcomes.

Cardiac ischemia and related outcomes.

One published study has examined evidence for the relationship of particulate air pollutant exposure to cardiac ischemia in humans. An epidemiologic study of 45 adults with stable CHD conducted by Pekkanen et al. (2002) analyzed data from repeated biweekly in-clinic electrocardiographic (ECG) measurements during submaximal exercise testing and outdoor UFPs and fine particles measured at a central regional site of Helsinki, Finland (Table 1). They found significant associations between risk of ST segment depression and ambient PM2.5 mass, number concentrations of ultrafine mode particles 0.01–0.1 μm in diameter (NC0.01–0.1), and number concentrations of accumulation-mode particles 0.1–1.0 μm in diameter (NC0.1–1) (Table 1). Odds ratios (ORs) were around 3.0 for all particle metrics for an increase around their interquartile distribution. Smaller but significant associations were also found for the gases NO2 and CO, which were moderately correlated with the co-located particle measurements. The association with UFP number concentration was independent of PM2.5 mass concentration. It is surprising that associations for out-door ambient NC0.01–0.1 were as strong as for PM2.5, given the expectation that human exposure to UFPs is less consistently represented by central site PM monitoring than is exposure to PM2.5 monitoring, which shows much lower spatial variability than UFPs (reviewed by Pekkanen and Kulmala 2004; Sioutas et al. 2005).

Cardiorespiratory symptoms potentially related to cardiac ischemia were assessed by de Hartog et al. (2003) in elderly patients with CHD. The authors found that although chest pain was not associated with PM exposure, a 10 μg/m3 increase in ambient PM2.5 was associated with shortness of breath and avoidance of activities (Table 1).

A case-crossover study of 691 subjects from the Augsburg Myocardial Infarction Registry found a 2- to 3-fold increased risk of MI for time–activity diary reports of hours exposed to traffic, particularly for times spent in cars and public transportation in the hours leading up to cardiac symptom onset (Peters et al. 2004) (Table 1). No direct air pollutant measurements were available. However, as discussed in our companion article (Sioutas et al. 2005), exposures to UFPs can be magnitudes higher than background levels within vehicles and near busy highways, and to a much greater degree than larger particles. Accumulation-mode PM, volatile organic compounds, and gases such as CO could have also played a role in the findings of Peters et al. (2004).

Blood pressure.

Two studies showing associations between air pollution and blood pressure (BP) followed subjects with COPD (Brauer et al. 2001; Linn et al. 1999; Table 1). Linn et al. (1999) found that for only 120 total person-observation times in 30 subjects, an increase of 33 μg/m3 ambient PM10 (study mean) was associated with a 5.7 mm mercury (Hg) increase in systolic BP. In contrast, Brauer et al. (2001) found systolic BP was inversely but weakly associated with personal PM2.5 in a pooled regression analysis of 16 subjects with COPD monitored on 7 separate days. This association was not confounded by inverse associations with ambient CO. Inverse associations with ambient PM10 were larger but were confounded by CO. Another study examined 2,607 German adults younger than 65 years evaluated on two occasions 3 years apart and found a positive association of systolic BP with ambient concentrations of both total suspended particulates (TSP) and SO2 (Ibald-Mulli et al. 2001) (Table 1).

Ibald-Mulli et al. (2004) conducted one of the few panel studies to focus on the relationship between UFPs and BP (Table 1). They followed 131 adults with CHD in three European centers every 2 weeks for about 11 clinic visits. An increase of a 5-day average of 10,000/cm3 UFPs (PM0.01–0.1) was associated with small decrease in systolic BP (−0.72 mm Hg; p < 0.01) and diastolic BP (−0.70 mm Hg; p < 0.01). Comparably small associations were also found for CO, 1,000/cm3 accumulation-mode particles, and 10 μg/m3 PM2.5. The authors hypothesized that BP medications in these CHD patients might have blunted or modified the response to air pollution exposure. However, these results contrast those of a panel study by Zanobetti et al. (2004), who found that ambient 5-day average PM2.5 was positively associated with BP among 62 patients with preexisting heart disease, using data from 631 repeated visits for cardiac rehabilitation in Boston (Table 1).

Panel study results for PM2.5 can be compared with two experimental human studies (Brook et al. 2002; Gong et al. 2003; not shown in Table 1). Gong et al. (2003) studied the effects of PM2.5 concentrated ambient particles (CAPs) from Los Angeles air versus clean air on systolic BP in 12 healthy versus 12 asthmatic adults using a 2-hr rest–exercise exposure period in a chamber. CAPs are used to approximate the effects of “real-world” particles. They found inverse associations of PM2.5 CAPs with systolic BP in asthmatics, but positive associations in healthy subjects. Results from two small studies by Brauer et al. (2001) and Gong et al. (2003) with relatively good exposure data show that PM2.5 mass is inversely associated with BP in subjects with obstructive lung diseases. Brook et al. (2002) also studied the vascular effects of 150 μg/m3 PM2.5 CAPs from Toronto air, adding 120 ppb O3, in 25 healthy adults using a 2-hr exposure period in a chamber. They found a significant but small 0.1 mm decrease in brachial artery diameter by ultrasonography for the joint exposures versus filtered air but no change in BP, flow-mediated diameter (endothelium dependent), or nitro-glycerin-mediated dilatation (endothelium independent). A follow-up analysis showed that the organic and EC fractions of PM2.5 CAPs were significant determinants of the effects on brachial artery diameter, which is a more sensitive biomarker of effect than BP (Urch et al. 2004).

Potential mechanisms for the observed PM-associated increases in BP have been suggested to include an increase in sympathetic tone and/or the modulation of basal systemic vascular tone due to increased concentrations of a plasma peptide known as endothelin-1 (Ibald-Mulli et al. 2001). Endothelin-1 has multiple cardiovascular actions, including vasoconstriction, leading to maintenance of basal vascular tone and BP (Haynes and Webb 1998) and accentuating BP elevation in more severe, sodium-sensitive hypertension (Schiffrin 2001). It is directly associated with the severity of CHF and risk of subsequent cardiac death in CHF patients (Galatius-Jensen et al. 1996; Tsutamoto et al. 1995). Endothelin-1 is produced and cleared in the lung and is generated in response to the presence of ROS (free radicals) and their metabolites (Haynes and Webb 1998). This leaves open the possibility that pollutants could induce an excess production of endothelin-1. Supporting evidence is that urban particles have been shown to increase endothelin-1 in rats (Bouthillier et al. 1998). Effects of endothelin-1 are partly counterbalanced by vasodilatory influences of endothelial nitric oxide (NO; Vanhoutte 2000). Endothelial NO synthase produces NO, which traverses the extracellular space to induce smooth muscle relaxation in the vessel wall. One ROS that can be produced in the presence of certain pollutant components is superoxide, which can react with NO to form the potent oxidant peroxynitrite. Peroxynitrite is likely involved in lipid peroxidation (O’Donnell and Freeman 2001). Therefore, an additional potential mechanism whereby pollutant components can increase BP includes superoxide-mediated inhibition of the actions of NO in inducing vasodilatation.

Despite the above data on potential biologic mechanisms, reviewed epidemiologic studies have found both a decrease and increase in BP in relation to air pollutant exposures. This may be because of differences between subject populations, differences in the types of regional air pollutants, or possibly due to medications used or underlying pathology (healthy, COPD, asthma, CHD, etc.). There is also a lack of data in most studies on other influences on BP, namely, emotional states and physical activity, which could have sustained influence on nonambulatory BP measurements. The above factors could result in contrasting shifts in sympathetic and vagal tone in response to inhaled air pollutants, or contrasting shifts in the balance between mediators such as endothelin-1 and endothelial NO. The time course of exposure–response relationships is also ill-defined, particularly periods of exposure averaging times ranging from minutes to days. None of the epidemiologic studies used ambulatory BP monitoring to assess acute effects of real time changes in exposure. Ambulatory BP monitoring is more closely associated with end organ damage (heart, kidney, brain) than isolated systolic or diastolic BP readings taken in clinic offices (Mancia and Parati 2000).

Autonomic control of cardiac rhythm.

Heart rate variability (HRV) is a widely used noninvasive method to investigate cardiovascular autonomic control. Reduced HRV has been shown to be a predictor of increased mortality after MI (Kleiger et al. 1987; La Rovere et al. 1998) and has been related especially to sudden arrhythmic death (Hartikainen et al. 1996; Odemuyiwa et al. 1991). Fourier analysis of HRV can show the magnitude of variance in the heart’s rhythm across different frequency bands. Different autonomic influences on cardiovascular function (HR and BP) are reflected by different frequency bands. The high-frequency (HF) band (0.15–0.40 Hz) has been used to estimate cardiac vagal control and is linked to respiratory influences (Task Force 1996). Lower frequencies (0.04–0.15 Hz) are believed to represent mixed sympathetic and parasympathetic influences (Task Force 1996). Time domain measurements are also used (described below).

One controlled exposure study showed significant decreases in HRV in 10 healthy elderly adults for 2-hr exposures to CAPs from Chapel Hill, North Carolina (mostly mobile source) compared with clean air, and the decrease persisted 24 hr later (Devlin et al. 2003). In epidemiologic studies discussed below, ambient PM has been associated with decreased HRV (Chan et al. 2004; Creason et al. 2001; Gold et al. 2000; Holguin et al. 2003; Liao et al. 1999; Magari 2002, Magari et al. 2001, 2002; Peters et al. 1999; Pope et al. 2004b, 1999) and cardiac arrhythmia (Peters et al. 2000). Only two studies to our knowledge have investigated effects of personal PM exposures on HRV (Chan et al. 2004; Magari et al. 2001), and one on personal CO (Tarkiainen et al. 2003).

Liao et al. (1999) showed that the largest inverse associations between nonambulatory HRV measures and PM2.5 were for subjects with a history of cardiovascular conditions, although the number subjects (18) was small and the specific illnesses were not separated (not shown in Table 1). Another study of 56 elderly subjects showed inverse associations of nonambulatory high- and low-frequency HRV with indoor and outdoor 24-hr gravimetric PM2.5 collected in a retirement home (Creason et al. 2001; not shown in Table 1). Using hourly ambient PM2.5 data, they briefly reported that models using prior 4-hr average PM2.5 and time-lagged 4-hr PM2.5 were similar in magnitude to effects of the 24-hr PM2.5 averages, suggesting a mixture of short-term and cumulative effects. Holguin et al. (2003) studied 34 elderly nursing home residents living in Mexico City and showed a strong decrease in the high-frequency component of HRV with high ambient PM2.5 exposure, and the association was stronger for indoor home PM2.5. Those with hypertension had the largest reductions in HRV (Table 1). Pope et al. (1999) also used ambulatory HR monitoring in 7 elderly subjects with respiratory and cardiovascular disease before, during, and after episodes of elevated pollution. They found that ambient PM10 was associated with decreased in the standard deviation (SD) of normal-to-normal (NN) intervals (SDNN), a time domain measure of overall HRV. However, they also found an increase in the square root of the mean of squared differences between adjacent NN intervals (r-MSSD; time domain measurement that corresponds to high-frequency variability and parasympathetic tone). A larger study using ambulatory ECG monitors by Pope et al. (2004b) found that ambient PM2.5 was associated with a decrease in both SDNN and r-MSSD in 88 elderly subjects in Utah (Table 1). Magari et al. (2001) studied 40 workers occupationally exposed to welding fumes and residual oil fly ash with 24-hr monitoring using ambulatory HR monitors and personal real-time PM2.5 measurements from a TSI Inc. DustTrak (Shoreview, MN) (Table 1). They found significant decreases in SD of average 5-min NN intervals in relation to increases in prior 1-hr moving averages of PM2.5. They also found increasingly greater decreases in SDNN for higher PM2.5 across longer PM2.5 averaging times up to 9 hr. Magari et al. (2001) suggested inhaled particles directly affect autonomic function through a sympathetic stress response, represented by their acute response finding, and/or secondarily through airway inflammation and cytokine release into the circulation, represented by their cumulative response finding. Riediker et al. (2004) placed portable air-quality monitors in patrol cars of nine healthy male North Carolina Highway Patrol troopers who wore ambulatory ECG monitors (Table 1). In-vehicle PM2.5 was positively associated with ectopic beats, heart beat cycle length, HF HRV, and SDNN.

Chan et al. (2004) conducted the only study to date to assess the relationship between HRV and particle number concentrations (dominated by UFPs) for particles 0.02–1.0 μm in diameter (NC0.02–1) (Table 1). They followed 9 young healthy adults (2 females) and 10 elderly male subjects with obstructive lung function impairment. This was also the first study to examine the effects of personal exposure to UFPs on HRV. Subjects were monitored over only 10 daytime hours using a P-Trak Ultrafine Particle Counter (TSI Inc.) for NC0.02–1. Subjects also wore ambulatory ECG monitors for continuous 5-min beat-to-beat intervals to assess HRV. Using linear mixed-effects models, they found that decreases in HRV indices (SDNN and r-MSSD) were associated with exposure to 1- to 4-hr moving averages of NC0.02–1 before the 5-min HRV measurements, adjusting for age, sex, body mass index, environmental tobacco smoke exposure, and temperature (Table 1). Associations were stronger for the elderly panel, with the strongest effects from 2-hr average NC0.02–1. These results along with those of Magari et al. (2001) suggest that the effect of personal PM exposure on autonomic function is acute, although the monitoring period (10 hr) was too short in the Chan et al. (2004) study to assess longer-term effects.

Tarkiainen et al. (2003) studied six patients with CHD for 1 day per week for 3 weeks with continuous personal CO exposure monitors, ambulatory ECG monitoring for HRV, and time–activity diaries and found r-MSSD increased in relation to high CO exposures (> 2.7 ppm peaks lasting 17 min, SD 8 min) (Table 1). This result contrasted results of most studies using PM exposures, except the study of Pope et al. (1999). No particle data were available, but it is again important to note that outdoor CO at sites close to dense traffic is highly correlated with UFPs (Zhu et al. 2002). It is conceivable that CO and/or UFPs increase vagal control and induce bradyarrhythmias.

In a study of arrhythmias and air pollution, investigators followed 100 subjects in eastern Massachusetts with implanted defibrillators (Peters et al. 2000; Table 1). They found that patients with 10 or more defibrillator discharge interventions for cardiac arrhythmias experienced increased arrhythmias in association with outdoor ambient NO2, CO, and black carbon, but PM2.5 was less strongly related. The most robust association was found for NO2, which may have been a marker for local traffic-related pollution, whereas particle mass may have been additionally influenced by other sources. Exposure was represented by only one Boston monitoring site.

Systemic inflammation and thrombosis.

The view that air-pollution–induced airway inflammation triggers systemic hypercoagulability (Seaton et al. 1995) has been supported in recent epidemiologic studies. It is relevant in this regard that, compared with unaffected people, patients with CHD (Lagrand et al. 1999; Mendall et al. 1997; Stec et al. 2000; Woods et al. 2000) or a complication of CHD, CHF (Pye et al. 1990; Torre-Amione et al. 1996), have increased levels of inflammatory cytokines such as interleukin (IL)-1β and IL-6, and tumor necrosis factor-α (TNF-α). They also have increased levels of circulating acute phase proteins such as C-reactive protein (CRP) and fibrinogen. In patients with CHD, CRP is also a strong independent predictor of future coronary events (Rifai and Ridker 2001). Cohort studies have shown that levels of acute phase proteins, cytokines, and hemostatic factors indicative of a thrombophilic state or endothelial activation are elevated at baseline in subjects at risk for future coronary occlusion or cardiovascular mortality (Cushman et al. 1999; Danesh et al. 2000; Folsom et al. 2001; Harris et al. 1999; Haverkate et al. 1997; Jager et al. 1999; Kuller et al. 1996; Lind et al. 2001; Malik et al. 2001; Ridker 2001; Ridker et al. 2000, 2001; Thompson et al. 1995). Air pollutant exposures that lead to acute increases in already elevated levels of inflammatory and hemostatic factors may also precipitate adverse health outcomes. This is a strong possibility in patients with diagnosed or underlying CHD, a population most likely driving the time-series associations. In addition, high air pollutant exposures that lead to chronic or repeated increases in systemic inflammation through oxidative stress responses to ROS may promote the progression of atherosclerosis in susceptible individuals.

Recent studies have shown acute associations between air pollutant exposures and systemic responses indicating inflammation and hypercoagulability. Seaton et al. (1999) studied 112 elderly individuals and used 1 day of personal PM10 data per person to predict the remaining 2 days using ambient (city center) PM10 data (Table 1). Results showed inverse associations of estimated personal PM10 with albumin-adjusted hemoglobin, packed cell volume, red blood cell count, platelets, and factor VII levels. They found no associations between PM10 and IL-6 or white blood cell count. Only ambient PM10 was positively associated with CRP concentrations, but it was also inversely associated with fibrinogen. The authors hypothesized that particles enter lung endothelial cells or erythrocytes and subsequently influence red cell adhesiveness, leading to peripheral sequestration of red cells. Contrasting results were found by Schwartz (2001), who used health data from the Third National Health and Nutrition Examination Survey (NHANES III) in the United States (Table 1). Results showed that outdoor PM10 levels on the day of subject visits or previous day was positively associated with fibrinogen levels and counts of platelets and white blood cells. Fibrinogen increased by 13 μg/dL (95% CI, 4.6–22.1) for an interquartile range change in PM10 of 26 μg/m3. PM effects were independent of gaseous pollutants. Schwartz (2001) argued that the NHANES III results were consistent with data in controlled human exposure (Ghio et al. 2000) and animal studies (Gardner et al. 2000) that showed increased plasma fibrinogen after particle exposures. Pekkanen et al. (2000) found no association between PM10 and fibrinogen using cross-sectional data from another cohort study of 7,205 subjects in London. However, they did find associations between fibrinogen and two pollutant gases, NO2 and CO, but not SO2 or O3. Epidemiologic studies in Augsburg, Germany, have also shown positive associations of ambient air pollution with plasma viscosity (Peters et al. 1997) and with CRP concentrations (Peters et al. 2001b) (Table 1). Another study of people exposed to forest fire smoke showed increased circulating levels of IL-1β and IL-6 (Van Eeden et al. 2001; not shown). A panel study by Pope et al. (2004b) (Table 1) with 88 elderly subjects in Utah showed a 0.81 mg/dL CRP increase in association with a 100 μg/m3 increase in ambient PM2.5. There was no association with white or red blood cell counts, platelets, or whole-blood viscosity. Riediker et al. (2004; discussed above) assessed the relationship between in-vehicle PM exposure and markers of inflammation in nine healthy male state troopers. An in-vehicle 10 μg/m3 PM2.5 increase was associated with decreased lymphocytes (−11%), increased red blood cell indices (1%), neutrophils (6%), CRP (32%), and von Willebrand factor (12%).

Summary and biologic plausibility.

To date only three studies have directly evaluated the effects on cardiovascular health by UFPs or particle number concentration (Chan et al. 2004; Ibald-Mulli et al. 2004; Pekkanen et al. 2002). Results of Pekkanen et al. (2002) showing ST segment depression in relation to UFPs are the most compelling findings. Associations of ambient NC0.01–0.1 with ST segment depression were independent of ambient PM2.5, but it is unclear whether the ambient exposure data represented personal UFP exposures of subjects. Other indirect evidence that components of fossil fuel combustion are important comes from studies using surrogate measures of particle composition such as black smoke, proximity of homes to traffic, or source apportionment data. Epidemiologic associations for pollutant gases also seem to support the idea that cardiovascular effects may be linked to primary products of combustion emissions that include UFPs.

Because hypertension, ST segment depression, and cardiac arrhythmias are well-known risk factors for cardiac morbidity and mortality, the above findings of acute associations with PM from individual-level studies are relevant to the reported findings of time-series and cohort investigations of mortality and hospital morbidity. However, mixed findings for BP have not provided a coherent view of particle effects. Findings for HRV are largely consistent in finding a decrease in HRV except for the increase in r-MSSD with ambient PM among elderly subjects found by Pope et al. (1999) and increased HF HRV for in-vehicle PM among healthy men found by Riediker et al. (2004). The clinical importance of HRV to cardiovascular disease is unclear however (Task Force 1996), and many technical issues regarding the influence of respiratory patterns (respiratory sinus arrhythmia) and psychosocial stress (both unmeasured in the reviewed studies) remain unresolved (Sloan et al. 1994).

The reviewed epidemiologic studies on circulating biomarkers of effect show inconsistent relationships between air pollution and blood markers of inflammation and hyper-coagulability, possibly because all but two studies used ambient exposure to PM. Currently, only the studies of Seaton et al. (1999) and Riediker et al. (2004) used any personal PM exposure measurements, but results are not consistent. In addition, the reviewed studies of circulating biomarkers did not target people with cardiovascular diseases, who are expected to be among the most susceptible population, as indicated in the time-series investigations.

The main limitation of most epidemiologic studies is exposure misclassification from dependence on central site rather than on personal or microenvironmental exposure data. However, studies reported above that do have personal exposure data also have limited numbers of subjects or days monitored. In general some major methodologic issues that remain involve choice of susceptible populations, personal exposure assessment, and timing of measurements to assess the temporality of exposure–dose–response relationships.

Despite the inconsistencies in epidemiologic data, sound postulated mechanisms support the biologic plausibility of many of the findings. Airway inflammation from PM likely involves inhalation of agents leading to the deposition or production in lung tissue of ROSs. The ROSs then induce subsequent oxidant injury and inflammatory responses (Pritchard et al. 1996; Schreck et al. 1991) both in the lungs and systemically. Inhalation of particle-bound airborne transition metals (copper, iron, nickel, vanadium) can lead to the production of ROSs in lung tissue. Residual oil fly ash containing high concentrations of transition metals but low in organic compounds have been shown to induce in vitro increases in IL-6 mRNA in human epithelial cells (Quay et al. 1998). Dogs exposed to CAPs from Boston air showed increased bronchoalveolar lavage macrophages and increased circulating neutrophils in relation to a vanadium/nickel factor, but no associations were shown with total mass (Clarke et al. 2000). This suggests that pollutant composition was important.

Organic constituents of PM are also capable of generating ROS. Nel et al. (2001) have presented evidence that polycyclic aromatic hydrocarbons (PAHs) from diesel exhaust particles (DEPs) and oxidized derivatives of PAHs, such as quinones, lead to the generation of ROSs and subsequent oxidant injury and inflammatory responses, including the production of nuclear transcription factor κB (NF-κB). NF-κB increases the transcription of cytokines and acute phase proteins (Schreck et al. 1991). Evidence has been presented that DEPs induce a broad polyclonal activation of cytokines from an adjuvant-like activity of DEP PAHs (Diaz-Sanchez et al. 1996, 1997; Fujieda et al. 1998; Nel et al. 1998, 2001). Human pulmonary responses to DEPs include increased neutrophils and B-lymphocytes in lavage fluids, increased expression of endothelial adhesion molecules ICAM-1 (intercellular adhesion molecule-1) and VCAM-1 (vascular cell adhesion molecule-1) in bronchial biopsies, and increased neutrophils and platelets in peripheral blood (Salvi et al. 1999). Such DEP-induced effects from oxidative stress mechanisms would be expected to lead to increased systemic hypercoagulability, but to date supporting data in humans are limited.

Epidemiologic evidence in humans that PM exposure increases biomarkers of oxidative stress in blood is limited to one study of 50 healthy young adults in Copenhagen using air samplers carried by subjects (Sorensen et al. 2003). They found a positive association between personal black carbon exposure and 2-aminoadipic semialdehyde in plasma proteins, a protein oxidation product. However, no association with personal PM2.5 mass was found, suggesting that traffic-related causal components may have been better represented by black carbon than by particle mass. A lipid peroxidation product (malondialdehyde), as well as red blood cell counts and hemoglobin concentrations, was positively associated with PM2.5 exposure in women only.

There are also plausible linkages between pulmonary and cardiovascular responses to PM. Airway inflammatory responses have been demonstrated in animals exposed to particulate air pollutants (U.S. EPA 2003). As discussed above, there is growing evidence that airway responses may trigger systemic inflammation and hypercoagulability. In addition, PM can induce neurogenic inflammation in the lungs from activation of capsaicin-sensitive irritant receptors, leading to the release of tachykinins from sensory terminals and then airway inflammation and bronchoconstriction (Veronesi and Oortgiesen 2001). This response could then affect cardiovascular autonomic function (Carr and Undem 2001; Yeates 2000), but it is not yet clear to what extent these mechanisms explain epidemiologic findings of air pollutant associations with cardiac rhythm and BP. There is limited evidence for an effect of tachykinins on cardiac function (Maggi 1996). In addition, the linkage between airway inflammation, cytokine/chemokine release, and autonomic stress response has not been directly demonstrated in humans. There are some in vitro data linking actions of pro-inflammatory cytokines IL-1β and TNF-α to myocardial cell changes in contractility and action potentials (DeMeules et al. 1992; Finkel et al. 1992; Li and Rozanski 1993; Yokoyama et al. 1993) and to induction of arrhythmias (Weisensee et al. 1993).

There are experimental data indirectly supporting a linkage between cellular inflammation in the lungs and cardiovascular responses to air pollutants. An experiment in hyperlipidemic rabbits showed that intrapharyngeal instillation of ambient urban PM10 led to an increase in circulating polymorphonuclear neutrophils and caused an increase in the volume fraction of atherosclerotic lesions, which correlated with the number of alveolar macrophages that phagocytosed PM10 in the lung (r = 0.5) (Suwa et al. 2002). Particle-induced airway inflammation and translocation of UFPs and other pollutants into the circulation could lead to an increase in thrombogenic and inflammatory activity in the blood and to a disturbance in cardiovascular function. These extrapulmonary effects are expected to increase the risk of adverse cardiovascular outcomes such as hospitalization.

Other evidence links airway inflammation with cardiovascular effects. Cohort data have shown links of COPD with CHD risk independent of other risk factors (Jousilahti et al. 1999; Wedzicha et al. 2000), suggesting that pulmonary inflammatory processes may have pro-inflammatory effects on the vascular endothelium. This could occur in individuals with asthma or COPD who have depleted antioxidant defenses from oxidative stress compared with normal subjects, and their defenses are further lowered during disease exacerbations (Rahman et al. 1996). Zanobetti et al. (2000a) have shown that a positive association between hospital admissions for cardiovascular diseases and ambient air pollution was nearly doubled in elderly patients admitted with concurrent respiratory infections. Diabetics appear to be another susceptible group, with stronger associations between cardiovascular hospital admissions and ambient air pollution (Zanobetti and Schwartz 2001).

Several excellent reviews of experimental data examining acute pulmonary and cardiovascular responses to inhaled UFPs and fine particles have proposed pathophysiologic mechanisms (American Thoracic Society 1999; Dhalla et al. 2000; Donaldson et al. 2001; Godleski et al. 2000; MacNee and Donaldson 2000; Nel et al. 2001; Utell and Frampton 2000; Utell et al. 2002; van Eeden and Hogg 2002). We have synthesized these and other data into the following proposed sequence of events for UFPs that link pulmonary and cardiovascular end points (Figure 1). Most of these mechanisms likely also apply to larger PM size fractions, particularly soluble components of PM2.5, and retained nonsoluble particles in the lung that may stimulate the bone marrow to induce similar systemic responses (van Eeden and Hogg 2002):

Figure 1.

Figure 1

Hypothesized pathways leading to adverse cardiovascular health effects from exposure to UFPs.

  • UFP exposure is followed by high pulmonary deposition (Chalupa et al. 2004; Daigle et al. 2003; International Commission on Radiological Protection 1994). UFPs and associated air toxics translocate to the interstitium and gain entry into the circulation (Nemmar et al. 2002, 2004; Oberdörster et al. 2002).

  • Redoxactive components of PM lead to the production of ROSs in various cells in the lungs, blood, and vascular tissues.

  • This is followed by oxidative stress responses in pulmonary epithelium and pulmonary vascular endothelium and in extrapulmonary vascular endothelium, leading to the production of oxidized phospholipids (especially LDL), lipid peroxidation (e.g., 8-iso-prostaglandin F), reduced antioxidant capacity (e.g., increase in the ratio of oxidized to reduced glutathione), and the production of superoxide anions by endothelial NADPH oxidase, all of which likely contribute to atherogenesis. Genetic polymorphisms in key metabolic enzymes likely play a role in susceptibility.

  • Pulmonary and extrapulmonary peripheral vascular oxidative stress results in the activation and mobilization of mononuclear leukocytes and the expression of NF-κB, followed by increases in pro-inflammatory cytokines (e.g., IL-1β, IL-6, and TNF-α) and endothelial cell activation.

  • Emigration of inflammatory cells from blood to tissue sites involves up-regulation of adhesion molecules (VCAM-1, ICAM-1) on vascular endothelium and circulating leukocytes.

  • Increased release of cytokines by activated mononuclear cells in the lungs and in the blood leads to initiation of hepatic synthesis of acute phase proteins (e.g., CRP and fibrinogen).

  • A hypercoagulable state then occurs with platelet activation, hemostasis, and blood clot formation followed by fibrinolytic activity; this increases the risk of a coronary event. Cytokines may also have direct effects on cardiac function.

  • Endothelial cell activation also leads the expression of endothelin-1, which induces vasoconstriction, and increased systolic and diastolic BP, and the expression of extracellular superoxide dismutase (SOD). SOD catalyzes superoxide (O2) to H2O2, which lowers endothelial NO-induced vasodilation. Neuroinflammatory responses involving tachykinins and catecholamines may also affect cardiovascular autonomic tone.

  • The systemic inflammatory response also stimulates the bone marrow to release leukocytes and platelets, and polymorphonuclear leukocytes increasingly sequester in pulmonary capillaries to induce more inflammation.

Conclusion

As presented in this review, numerous studies have implicated particulate air pollution as an important contributor to morbidity and mortality from cardiovascular causes. Most of these data have been epidemiologic and have used available air pollution data from governmental monitoring stations. Because such data are collected to meet regulatory standards, they may not meet the needs of researchers trying to understand the causal pollutant components that lead to specific adverse health effects. UFPs and related toxic constituents and precursors are examples of air pollutants that have not been fully investigated, in part due to lack of available data. To date, data from epidemiologic studies indirectly implicate traffic- and other combustion-related pollutants, which include UFPs. Exposure assessment issues for UFPs are complex and need to be considered before undertaking epidemiologic investigations of UFP health effects (Sioutas et al. 2005).

A large body of evidence shows that inflammation and oxidative stress are related to both acute changes in cardiovascular health and chronic processes, including atherosclerosis. It is likely that redoxactive components in UFPs from fossil fuel combustion reach target sites in the lungs, vasculature, and heart to induce inflammation and oxidative stress, adding to the burden of known lifestyle risk factors for cardiovascular disease such as diet, tobacco smoke, and stress.

Footnotes

This work was supported by grant ES-12243 from the National Institute of Environmental Health Sciences (NIEHS), National Institutes of Health (NIH); the contents of this article are solely the responsibility of the author and do not necessarily represent the official views of the NIEHS, NIH. This work was also supported by the Southern California Particle Center and Supersite funded by the U.S. Environmental Protection Agency (U.S. EPA; STAR award R82735201).

This manuscript has not been subjected to the U.S. EPA peer and policy review and therefore does not necessarily reflect the views of the agencies. No official endorsement should be inferred.

References

  1. Abbey DE, Nishino N, McDonnell WF, Burchette RJ, Knutsen SF, Beeson WL, et al. Long-term inhalable particles and other air pollutants related to mortality in nonsmokers. Am J Respir Crit Care Med. 1999;159:373–382. doi: 10.1164/ajrccm.159.2.9806020. [DOI] [PubMed] [Google Scholar]
  2. American Thoracic Society. Recommendations for standardized procedures for the online and offline measurement of exhaled lower respiratory nitric oxide and nasal nitric oxide in adults and children. Am J Respir Crit Care Med. 1999;160:2104–2117. doi: 10.1164/ajrccm.160.6.ats8-99. [DOI] [PubMed] [Google Scholar]
  3. Bouthillier L, Vincent R, Goegan P, Adamson IY, Bjarnason S, Stewart M, et al. Acute effects of inhaled urban particles and ozone: lung morphology, macrophage activity, and plasma endothelin-1. Am J Pathol. 1998;153:1873–1884. doi: 10.1016/S0002-9440(10)65701-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Brauer M, Ebelt ST, Fisher TV, Brumm J, Petkau AJ, Vedal S. Exposure of chronic obstructive pulmonary disease patients to particles: respiratory and cardiovascular health effects. J Expo Anal Environ Epidemiol. 2001;11:490–500. doi: 10.1038/sj.jea.7500195. [DOI] [PubMed] [Google Scholar]
  5. Brook RD, Brook JR, Urch B, Vincent R, Rajagopalan S, Silverman F. Inhalation of fine particulate air pollution and ozone causes acute arterial vasoconstriction in healthy adults. Circulation. 2002;105:1534–1536. doi: 10.1161/01.cir.0000013838.94747.64. [DOI] [PubMed] [Google Scholar]
  6. Burnett RT, Cakmak S, Brook JR, Krewski D. The role of particulate size and chemistry in the association between summertime ambient air pollution and hospitalization for cardiorespiratory diseases. Environ Health Perspect. 1997a;105:614–620. doi: 10.1289/ehp.97105614. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Burnett RT, Dales R, Krewski D, Vincent R, Dann T, Brook JR. Associations between ambient particulate sulfate and admissions to Ontario hospitals for cardiac and respiratory diseases. Am J Epidemiol. 1995;142:15–22. doi: 10.1093/oxfordjournals.aje.a117540. [DOI] [PubMed] [Google Scholar]
  8. Burnett RT, Dales RE, Brook JR, Raizenne ME, Krewski D. Association between ambient carbon monoxide levels and hospitalizations for congestive heart failure in the elderly in 10 Canadian cities. Epidemiology. 1997b;8:162–167. doi: 10.1097/00001648-199703000-00007. [DOI] [PubMed] [Google Scholar]
  9. Burnett RT, Smith-Doiron M, Stieb D, Cakmak S, Brook JR. Effects of particulate and gaseous air pollution on cardiorespiratory hospitalizations. Arch Environ Health. 1999;54:130–139. doi: 10.1080/00039899909602248. [DOI] [PubMed] [Google Scholar]
  10. Carr MJ, Undem BJ. Inflammation-induced plasticity of the afferent innervation of the airways. Environ Health Perspect. 2001;109(suppl 4):567–571. doi: 10.1289/ehp.01109s4567. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Chalupa DC, Morrow PE, Oberdorster G, Utell MJ, Frampton MW. Ultrafine particle deposition in subjects with asthma. Environ Health Perspect. 2004;112:879–182. doi: 10.1289/ehp.6851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Chan C-C, Chuang KJ, Shiao GM, Lin LY. Personal exposure to submicrometer particles and heart rate variability in human subjects. Environ Health Perspect. 2004;112:1063–1067. doi: 10.1289/ehp.6897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Clancy L, Goodman P, Sinclair H, Dockery DW. Effect of air-pollution control on death rates in Dublin, Ireland: an intervention study. Lancet. 2002;360:1210–1214. doi: 10.1016/S0140-6736(02)11281-5. [DOI] [PubMed] [Google Scholar]
  14. Clarke RW, Coull B, Reinisch U, Catalano P, Killingsworth CR, Koutrakis P, et al. Inhaled concentrated ambient particles are associated with hematologic and bronchoalveolar lavage changes in canines. Environ Health Perspect. 2000;108:1179–1187. doi: 10.1289/ehp.001081179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Creason J, Neas L, Walsh D, Williams R, Sheldon L, Liao D, et al. Particulate matter and heart rate variability among elderly retirees: the Baltimore 1998 PM study. J Expo Anal Environ Epidemiol. 2001;11:116–122. doi: 10.1038/sj.jea.7500154. [DOI] [PubMed] [Google Scholar]
  16. Cushman M, Lemaitre RN, Kuller LH, Psaty BM, Macy EM, Sharrett AR, et al. Fibrinolytic activation markers predict myocardial infarction in the elderly. The Cardiovascular Health Study. Arterioscler Thromb Vasc Biol. 1999;19:493–438. doi: 10.1161/01.atv.19.3.493. [DOI] [PubMed] [Google Scholar]
  17. Daigle CC, Chalupa DC, Gibb FR, Morrow PE, Oberdorster G, Utell MJ, et al. Ultrafine particle deposition in humans during rest and exercise. Inhal Toxicol. 2003;15:539–552. doi: 10.1080/08958370304468. [DOI] [PubMed] [Google Scholar]
  18. Danesh J, Whincup P, Walker M, Lennon L, Thomson A, Appleby P, et al. Low grade inflammation and coronary heart disease: prospective study and updated meta-analyses. Br Med J. 2000;321:199–204. doi: 10.1136/bmj.321.7255.199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. de Hartog JJ, Hoek G, Peters A, Timonen KL, Ibald-Mulli A, Brunekreef B, et al. Effects of fine and ultrafine particles on cardiorespiratory symptoms in elderly subjects with coronary heart disease: the ULTRA study. Am J Epidemiol. 2003;157:613–623. doi: 10.1093/aje/kwg021. [DOI] [PubMed] [Google Scholar]
  20. DeMeules JE, Pigula FA, Mueller M, Raymond SJ, Gamelli RL. Tumor necrosis factor and cardiac function. J Trauma. 1992;32:686–692. doi: 10.1097/00005373-199206000-00003. [DOI] [PubMed] [Google Scholar]
  21. Desai MM, Zhang P, Hennessy CH. Surveillance for morbidity and mortality among older adults—United States, 1995–1996. Mor Mortal Wkly Rep CDC Surveill Summ. 1999;48:7–25. [PubMed] [Google Scholar]
  22. Devlin RB, Ghio AJ, Kehrl H, Sanders G, Cascio W. Elderly humans exposed to concentrated air pollution particles have decreased heart rate variability. Eur Respir J. 2003;40(suppl):76s–80s. doi: 10.1183/09031936.03.00402403. [DOI] [PubMed] [Google Scholar]
  23. Dhalla NS, Temsah RM, Netticadan T. Role of oxidative stress in cardiovascular diseases. J Hypertens. 2000;18:655–673. doi: 10.1097/00004872-200018060-00002. [DOI] [PubMed] [Google Scholar]
  24. Diaz-Sanchez D, Tsien A, Casillas A, Dotson AR, Saxon A. Enhanced nasal cytokine production in human beings after in vivo challenge with diesel exhaust particles. J Allergy Clin Immunol. 1996;98:114–123. doi: 10.1016/s0091-6749(96)70233-6. [DOI] [PubMed] [Google Scholar]
  25. Diaz-Sanchez D, Tsien A, Fleming J, Saxon A. Combined diesel exhaust particulate and ragweed allergen challenge markedly enhances human in vivo nasal ragweed-specific IgE and skews cytokine production to a T helper cell 2-type pattern. J Immunol. 1997;158:2406–2413. [PubMed] [Google Scholar]
  26. D’Ippoliti D, Forastiere F, Ancona C, Agabiti N, Fusco D, Michelozzi P, et al. Air pollution and myocardial infarction in Rome: a case-crossover analysis. Epidemiology. 2003;14:528–535. doi: 10.1097/01.ede.0000082046.22919.72. [DOI] [PubMed] [Google Scholar]
  27. Dockery DW, Pope CA, Xu X, Spengler JD, Ware JH, Fay ME, et al. An association between air pollution and mortality in six U.S. cities. N Engl J Med. 1993;329:1753–1759. doi: 10.1056/NEJM199312093292401. [DOI] [PubMed] [Google Scholar]
  28. Dominici F, McDermott A, Daniels M, Zeger SL, Samet JM. 2003. Revised analysis of the National Morbidity, Mortality, and Air Pollution Study. Part II: Mortality among residents of 90 cities. In: Health Effects Institute Research: Revised Analysis of Time Series Studies of Air Pollution and Health. Boston:Health Effects Institute, 9–24.
  29. Donaldson K, Stone V, Seaton A, MacNee W. Ambient particle inhalation and the cardiovascular system: potential mechanisms. Environ Health Perspect. 2001;109(suppl 4):523–527. doi: 10.1289/ehp.01109s4523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Finkel MS, Oddis CV, Jacob TD, Watkins SC, Hattler BG, Simmons RL. Negative inotropic effects of cytokines on the heart mediated by nitric oxide. Science. 1992;257:387–389. doi: 10.1126/science.1631560. [DOI] [PubMed] [Google Scholar]
  31. Folsom AR, Aleksic N, Park E, Salomaa V, Juneja H, Wu KK. Prospective study of fibrinolytic factors and incident coronary heart disease: the Atherosclerosis Risk in Communities (ARIC) Study. Arterioscler Thromb Vasc Biol. 2001;21:611–617. doi: 10.1161/01.atv.21.4.611. [DOI] [PubMed] [Google Scholar]
  32. Fujieda S, Diaz-Sanchez D, Saxon A. Combined nasal challenge with diesel exhaust particles and allergen induces in vivo IgE isotype switching. Am J Respir Cell Mol Biol. 1998;19:507–512. doi: 10.1165/ajrcmb.19.3.3143. [DOI] [PubMed] [Google Scholar]
  33. Galatius-Jensen S, Wroblewski H, Emmeluth C, Bie P, Haunso S, Kastrup J. Plasma endothelin in congestive heart failure: a predictor of cardiac death? J Card Fail. 1996;2:71–76. doi: 10.1016/s1071-9164(96)80025-x. [DOI] [PubMed] [Google Scholar]
  34. Gardner SY, Lehmann JR, Costa DL. Oil fly ash-induced elevation of plasma fibrinogen levels in rats. Toxicol Sci. 2000;56:175–180. doi: 10.1093/toxsci/56.1.175. [DOI] [PubMed] [Google Scholar]
  35. Ghio AJ, Kim C, Devlin RB. Concentrated ambient air particles induce mild pulmonary inflammation in healthy human volunteers. Am J Respir Crit Care Med. 2000;162:981–988. doi: 10.1164/ajrccm.162.3.9911115. [DOI] [PubMed] [Google Scholar]
  36. Godleski JJ, Verrier RL, Koutrakis P, Catalano P, Coull B, Reinisch U, et al. Mechanisms of morbidity and mortality from exposure to ambient air particles. Res Rep Health Eff Inst. 2000;91:5–88. [PubMed] [Google Scholar]
  37. Gold DR, Litonjua A, Schwartz J, Lovett E, Larson A, Nearing B, et al. Ambient pollution and heart rate variability. Circulation. 2000;101:1267–1273. doi: 10.1161/01.cir.101.11.1267. [DOI] [PubMed] [Google Scholar]
  38. Goldberg MS, Burnett RT, Bailar JC, III, Tamblyn R, Ernst P, Flegel K, et al. Identification of persons with cardio-respiratory conditions who are at risk of dying from the acute effects of ambient air particles. Environ Health Perspect. 2001a;109(suppl 4):487–494. doi: 10.1289/ehp.01109s4487. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Goldberg MS, Burnett RT, Brook J, Bailar JC, III, Valois MF, Vincent R. Associations between daily cause-specific mortality and concentrations of ground-level ozone in Montreal, Quebec. Am J Epidemiol. 2001b;154:817–826. doi: 10.1093/aje/154.9.817. [DOI] [PubMed] [Google Scholar]
  40. Gong H, Jr, Linn WS, Sioutas C, Terrell SL, Clark KW, Anderson KR, et al. Controlled exposures of healthy and asthmatic volunteers to concentrated ambient fine particles in Los Angeles. Inhal Toxicol. 2003;15:305–325. doi: 10.1080/08958370304455. [DOI] [PubMed] [Google Scholar]
  41. Harris TB, Ferrucci L, Tracy RP, Corti MC, Wacholder S, Ettinger WH, Jr, et al. Associations of elevated interleukin-6 and C-reactive protein levels with mortality in the elderly. Am J Med. 1999;106:506–512. doi: 10.1016/s0002-9343(99)00066-2. [DOI] [PubMed] [Google Scholar]
  42. Hartikainen JEK, Malik M, Stauton A, Poloniecki J, Camm J. Distinction between arrhythmic and nonarrhythmic death after acute myocardial infarction based on heart rate variability, signal-averaged electrocardiogram, ventricular arrhythmias and left ventricular ejection fraction. J Am Coll Cardiol. 1996;28:296–304. doi: 10.1016/0735-1097(96)00169-6. [DOI] [PubMed] [Google Scholar]
  43. Haverkate F, Thompson SG, Pyke SD, Gallimore JR, Pepys MB. Production of C-reactive protein and risk of coronary events in stable and unstable angina. European Concerted Action on Thrombosis and Disabilities Angina Pectoris Study Group. Lancet. 1997;349:462–466. doi: 10.1016/s0140-6736(96)07591-5. [DOI] [PubMed] [Google Scholar]
  44. Haynes WG, Webb DJ. Endothelin as a regulator of cardiovascular function in health and disease. J Hypertens. 1998;16:1081–1098. doi: 10.1097/00004872-199816080-00001. [DOI] [PubMed] [Google Scholar]
  45. Hoek G, Brunekreef B, Fischer P, van Wijnen J. The association between air pollution and heart failure, arrhythmia, embolism, thrombosis, and other cardiovascular causes of death in a time series study. Epidemiology. 2001;12:355–357. doi: 10.1097/00001648-200105000-00017. [DOI] [PubMed] [Google Scholar]
  46. Hoek G, Brunekreef B, Goldbohm S, Fischer P, van den Brandt PA. Association between mortality and indicators of traffic-related air pollution in the Netherlands: a cohort study. Lancet. 2002;360:1203–1209. doi: 10.1016/S0140-6736(02)11280-3. [DOI] [PubMed] [Google Scholar]
  47. Holguin F, Tellez-Rojo MM, Hernandez M, Cortez M, Chow JC, Watson JG, et al. Air pollution and heart rate variability among the elderly in Mexico City. Epidemiology. 2003;14:521–527. doi: 10.1097/01.ede.0000081999.15060.ae. [DOI] [PubMed] [Google Scholar]
  48. Ibald-Mulli A, Stieber J, Wichmann HE, Koenig W, Peters A. Effects of air pollution on blood pressure: a population-based approach. Am J Public Health. 2001;91:571–577. doi: 10.2105/ajph.91.4.571. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Ibald-Mulli A, Timonen KL, Peters A, Heinrich J, Wolke G, Lanki T, et al. Effects of particulate air pollution on blood pressure and heart rate in subjects with cardiovascular disease: a multicenter approach. Environ Health Perspect. 2004;112:369–377. doi: 10.1289/ehp.6523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. International Commission on Radiological Protection 1994. Human Respiratory Tract Model for Radiological Protection. Tarrytown, NY:Elsevier Science Ltd.
  51. Jager A, van Hinsbergh VW, Kostense PJ, Emeis JJ, Yudkin JS, Nijpels G, et al. von Willebrand factor, C-reactive protein, and 5-year mortality in diabetic and nondiabetic subjects: the Hoorn Study. Arterioscler Thromb Vasc Biol. 1999;19:3071–3078. doi: 10.1161/01.atv.19.12.3071. [DOI] [PubMed] [Google Scholar]
  52. Janssen NA, Schwartz J, Zanobetti A, Suh HH. Air conditioning and source-specific particles as modifiers of the effect of PM10 on hospital admissions for heart and lung disease. Environ Health Perspect. 2002;110:43–49. doi: 10.1289/ehp.0211043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Jousilahti P, Salomaa V, Rasi V, Vahtera E. Symptoms of chronic bronchitis, haemostatic factors, and coronary heart disease risk. Atherosclerosis. 1999;142:403–407. doi: 10.1016/s0021-9150(98)00248-2. [DOI] [PubMed] [Google Scholar]
  54. Katsouyanni K, Touloumi G, Samoli E, Gryparis A, Le Tertre A, Monopolis Y, et al. Confounding and effect modification in the short-term effects of ambient particles on total mortality: results from 29 European cities within the APHEA2 project. Epidemiology. 2001;12:521–531. doi: 10.1097/00001648-200109000-00011. [DOI] [PubMed] [Google Scholar]
  55. Kim S, Shen S, Sioutas C. Size distribution and diurnal and seasonal trends of ultrafine particles in source and receptor sites of the Los Angeles basin. J Air Waste Manage Assoc. 2002;52:297–307. doi: 10.1080/10473289.2002.10470781. [DOI] [PubMed] [Google Scholar]
  56. Kittelson DB. Engines and nanoparticles: a review. J Aerosol Sci. 1998;29:575–588. [Google Scholar]
  57. Kleiger RE, Miller JP, Bigger JT, Jr, Moss AJ. Decreased heart rate variability and its association with increased mortality after acute myocardial infarction. Am J Cardiol. 1987;59:256–262. doi: 10.1016/0002-9149(87)90795-8. [DOI] [PubMed] [Google Scholar]
  58. Koken PJM, Piver WT, Ye F, Elixhauser A, Olsen LM, Portier CJ, et al. Temperature, air pollution, and hospitalization for cardiovascular diseases among elderly people in Denver. Environ Health Perspect. 2003;111:1312–1317. doi: 10.1289/ehp.5957. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Kuller LH, Tracy RP, Shaten J, Meilahn EN. Relation of C-reactive protein and coronary heart disease in the MRFIT nested case-control study. Multiple Risk Factor Intervention Trial. Am J Epidemiol. 1996;144:537–547. doi: 10.1093/oxfordjournals.aje.a008963. [DOI] [PubMed] [Google Scholar]
  60. Künzli N, Jerrett M, Mack WJ, Beckerman B, LaBree L, Gilliland F, et al. Ambient air pollution and atherosclerosis in Los Angeles. Environ Health Perspect. 2004;113(2):201–206. doi: 10.1289/ehp.7523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Kwon HJ, Cho SH, Nyberg F, Pershagen G. Effects of ambient air pollution on daily mortality in a cohort of patients with congestive heart failure. Epidemiology. 2001;12:413–419. doi: 10.1097/00001648-200107000-00011. [DOI] [PubMed] [Google Scholar]
  62. La Rovere MT, Bigger JT, Jr, Marcus FI, Mortara A, Schwartz PJ. Baroreflex sensitivity and heart-rate variability in prediction of total cardiac mortality after myocardial infarction. ATRAMI (Autonomic Tone and Reflexes After Myocardial Infarction) Investigators. Lancet. 1998;351:478–484. doi: 10.1016/s0140-6736(97)11144-8. [DOI] [PubMed] [Google Scholar]
  63. Laden F, Neas LM, Dockery DW, Schwartz J. Association of fine particulate matter from different sources with daily mortality in six U.S. cities. Environ Health Perspect. 2000;108:941–947. doi: 10.1289/ehp.00108941. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Lagrand WK, Visser CA, Hermens WT, Niessen HW, Verheugt FW, Wolbink GJ, et al. C-reactive protein as a cardiovascular risk factor: more than an epiphenomenon? Circulation. 1999;100:96–102. doi: 10.1161/01.cir.100.1.96. [DOI] [PubMed] [Google Scholar]
  65. Le Tertre A, Medina S, Samoli E, Forsberg B, Michelozzi P, Boumghar A, et al. Short-term effects of particulate air pollution on cardiovascular diseases in eight European cities. J Epidemiol Community Health. 2002;56:773–779. doi: 10.1136/jech.56.10.773. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Li YH, Rozanski GJ. Effects of human recombinant interleukin-1 on electrical properties of guinea pig ventricular cells. Cardiovasc Res. 1993;27:525–530. doi: 10.1093/cvr/27.3.525. [DOI] [PubMed] [Google Scholar]
  67. Liao D, Creason J, Shy C, Williams R, Watts R, Zweidinger R. Daily variation of particulate air pollution and poor cardiac autonomic control in the elderly. Environ Health Perspect. 1999;107:521–525. doi: 10.1289/ehp.99107521. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Lin CA, Amador Pereira LA, de Souza Conceicao GM, Kishi HS, Milani R, Jr, Ferreira Braga AL, et al. Association between air pollution and ischemic cardiovascular emergency room visits. Environ Res. 2003;92:57–63. doi: 10.1016/s0013-9351(02)00054-3. [DOI] [PubMed] [Google Scholar]
  69. Lind P, Hedblad B, Stavenow L, Janzon L, Eriksson KF, Lindgarde F. Influence of plasma fibrinogen levels on the incidence of myocardial infarction and death is modified by other inflammation-sensitive proteins: a long-term cohort study. Arterioscler Thromb Vasc Biol. 2001;21:452–458. doi: 10.1161/01.atv.21.3.452. [DOI] [PubMed] [Google Scholar]
  70. Linn WS, Gong H, Jr, Clark KW, Anderson KR. Day-to-day particulate exposures and health changes in Los Angeles area residents with severe lung disease. J Air Waste Manag Assoc. 1999;49(spec no 9 ):108–115. doi: 10.1080/10473289.1999.10463890. [DOI] [PubMed] [Google Scholar]
  71. Linn WS, Szlachcic Y, Gong H, Jr, Kinney PL, Berhane KT. Air pollution and daily hospital admissions in metropolitan Los Angeles. Environ Health Perspect. 2000;108:427–434. doi: 10.1289/ehp.00108427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. MacNee W, Donaldson K. How can ultrafine particles be responsible for increased mortality? Monaldi Arch Chest Dis. 2000;55:135–139. [PubMed] [Google Scholar]
  73. Magari SR. The association of particulate air metal concentrations with heart rate variability. Environ Health Perspect. 2002;110:875–880. doi: 10.1289/ehp.02110875. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Magari SR, Hauser R, Schwartz J, Williams PL, Smith TJ, Christiani DC. Association of heart rate variability with occupational and environmental exposure to particulate air pollution. Circulation. 2001;104:986–991. doi: 10.1161/hc3401.095038. [DOI] [PubMed] [Google Scholar]
  75. Magari SR, Schwartz J, Williams PL, Hauser R, Smith TJ, Christiani DC. The association between personal measurements of environmental exposure to particulates and heart rate variability. Epidemiology. 2002;13:305–310. doi: 10.1097/00001648-200205000-00011. [DOI] [PubMed] [Google Scholar]
  76. Maggi CA. Tachykinins in the autonomic nervous system. Pharmacol Res. 1996;33:161–170. doi: 10.1006/phrs.1996.0023. [DOI] [PubMed] [Google Scholar]
  77. Malik I, Danesh J, Whincup P, Bhatia V, Papacosta O, Walker M, et al. Soluble adhesion molecules and prediction of coronary heart disease: a prospective study and meta-analysis. Lancet. 2001;358:971–976. doi: 10.1016/S0140-6736(01)06104-9. [DOI] [PubMed] [Google Scholar]
  78. Mancia G, Parati G. Ambulatory blood pressure monitoring and organ damage. Hypertension. 2000;36:894–900. doi: 10.1161/01.hyp.36.5.894. [DOI] [PubMed] [Google Scholar]
  79. Mann JK, Tager IB, Lurmann F, Segal M, Quesenberry CP, Jr, Lugg MM, et al. Air pollution and hospital admissions for ischemic heart disease in persons with congestive heart failure or arrhythmia. Environ Health Perspect. 2002;110:1247–1252. doi: 10.1289/ehp.021101247. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. McGrath JJ. Biological plausibility for carbon monoxide as a copollutant in PM epidemiologic studies. Inhal Toxicol. 2000;12(suppl 4):91–107. doi: 10.1080/089583700750019521. [DOI] [PubMed] [Google Scholar]
  81. Mendall MA, Patel P, Asante M, Ballam L, Morris J, Strachan DP, et al. Relation of serum cytokine concentrations to cardiovascular risk factors and coronary heart disease. Heart. 1997;78:273–277. doi: 10.1136/hrt.78.3.273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Moolgavkar SH. Air pollution and daily mortality in three U.S. counties. Environ Health Perspect. 2000;108:777–784. doi: 10.1289/ehp.00108777. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Morris RD, Naumova EN. Carbon monoxide and hospital admissions for congestive heart failure: evidence of an increased effect at low temperatures. Environ Health Perspect. 1998;106:649–653. doi: 10.1289/ehp.98106649. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Morris RD, Naumova EN, Munasinghe RL. Ambient air pollution and hospitalization for congestive heart failure among elderly people in seven large US cities. Am J Public Health. 1995;85:1361–1365. doi: 10.2105/ajph.85.10.1361. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. NRC (National Research Council) Committee on Research Priorities for Airborne Particulate Matter. 1998. Research Priorities for Airborne Particulate Matter. I: Immediate Priorities and Long-Range Research Portfolio. Washington, DC:National Academy Press.
  86. NRC (National Research Council) Committee on Research Priorities for Airborne Particulate Matter. 1999. Research Priorities for Airborne Particulate Matter. II: Evaluating Research Progress and Evaluating the Portfolio. Washington, DC: National Academy Press.
  87. NRC (National Research Council) Committee on Research Priorities for Airborne Particulate Matter. 2001. Research Priorities for Airborne Particulate Matter. III: Early Research Progress. Washington, DC:National Academy Press.
  88. NRC (National Research Council) Committee on Research Priorities for Airborne Particulate Matter. 2004. Research Priorities for Airborne Particulate Matter. IV: Continuing Research Progress. Washington, DC:National Academy Press.
  89. Nel AE, Diaz-Sanchez D, Li N. The role of particulate pollutants in pulmonary inflammation and asthma: evidence for the involvement of organic chemicals and oxidative stress. Curr Opin Pulm Med. 2001;7:20–26. doi: 10.1097/00063198-200101000-00004. [DOI] [PubMed] [Google Scholar]
  90. Nel AE, Diaz-Sanchez D, Ng D, Hiura T, Saxon A. Enhancement of allergic inflammation by the interaction between diesel exhaust particles and the immune system. J Allergy Clin Immunol. 1998;102:539–554. doi: 10.1016/s0091-6749(98)70269-6. [DOI] [PubMed] [Google Scholar]
  91. Nemmar A, Hoet PH, Vanquickenborne B, Dinsdale D, Thomeer M, Hoylaerts MF, et al. Passage of inhaled particles into the blood circulation in humans. Circulation. 2002;105:411–414. doi: 10.1161/hc0402.104118. [DOI] [PubMed] [Google Scholar]
  92. Nemmar A, Hoylaerts MF, Hoet PH, Nemery B. Possible mechanisms of the cardiovascular effects of inhaled particles: systemic translocation and prothrombotic effects. Toxicol Lett. 2004;149:243–253. doi: 10.1016/j.toxlet.2003.12.061. [DOI] [PubMed] [Google Scholar]
  93. Oberdörster G. Pulmonary effects of inhaled ultrafine particles. Int Arch Occup Environ Health. 2001;74:1–8. doi: 10.1007/s004200000185. [DOI] [PubMed] [Google Scholar]
  94. Oberdörster G, Sharp Z, Atudorei V, Elder A, Gelein R, Lunts A, et al. Extrapulmonary translocation of ultrafine carbon particles following whole-body inhalation exposure of rats. J Toxicol Environ Health A. 2002;65:1531–1543. doi: 10.1080/00984100290071658. [DOI] [PubMed] [Google Scholar]
  95. Odenmuyiwa O, Malik M, Farrell T, Bashir Y, Poloniecki J, Camm J. Comparison of the predictive characteristics of heart rate variability index and left ventricular ejection fraction for all-cause mortality, arrhythmic events and sudden death after acute myocardial infarction. Am J Cardiol. 1991;68:434–439. doi: 10.1016/0002-9149(91)90774-f. [DOI] [PubMed] [Google Scholar]
  96. O’Donnell VB, Freeman BA. Interactions between nitric oxide and lipid oxidation pathways: implications for vascular disease. Circ Res. 2001;88:12–21. doi: 10.1161/01.res.88.1.12. [DOI] [PubMed] [Google Scholar]
  97. Pekkanen J, Brunner EJ, Anderson HR, Tiittanen P, Atkinson RW. Daily concentrations of air pollution and plasma fibrinogen in London. Occup Environ Med. 2000;57:818–822. doi: 10.1136/oem.57.12.818. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Pekkanen J, Kulmala M. Exposure assessment of ultra-fine particles in epidemiologic time-series studies. Scand J Work Environ Health. 2004;30 (suppl 2):9–18. [PubMed] [Google Scholar]
  99. Pekkanen J, Peters A, Hoek G, Tiittanen P, Brunekreef B, de Hartog J, et al. Particulate air pollution and risk of ST-segment depression during repeated submaximal exercise tests among subjects with coronary heart disease: the Exposure and Risk Assessment for Fine and Ultrafine Particles in Ambient Air (ULTRA) study. Circulation. 2002;106:933–938. doi: 10.1161/01.cir.0000027561.41736.3c. [DOI] [PubMed] [Google Scholar]
  100. Peters A, Dockery DW, Muller JE, Mittleman MA. Increased particulate air pollution and the triggering of myocardial infarction. Circulation. 2001a;103:2810–2815. doi: 10.1161/01.cir.103.23.2810. [DOI] [PubMed] [Google Scholar]
  101. Peters A, Doring A, Wichmann HE, Koenig W. Increased plasma viscosity during an air pollution episode: a link with mortality? Lancet. 1997;349:1582–1587. doi: 10.1016/S0140-6736(97)01211-7. [DOI] [PubMed] [Google Scholar]
  102. Peters A, Frohlich M, Doring A, Immervoll T, Wichmann HE, Hutchinson WL, et al. Particulate air pollution is associated with an acute phase response in men; results from the MONICA-Augsburg Study. Eur Heart J. 2001b;22:1198–1204. doi: 10.1053/euhj.2000.2483. [DOI] [PubMed] [Google Scholar]
  103. Peters A, Liu E, Verrier RL, Schwartz J, Gold DR, Mittleman M, et al. Air pollution and incidence of cardiac arrhythmia. Epidemiology. 2000;11:11–17. doi: 10.1097/00001648-200001000-00005. [DOI] [PubMed] [Google Scholar]
  104. Peters A, Perz S, Dööring A, Stieber J, Koenig W, Wichmann HE. Increases in heart rate during an air pollution episode. Am J Epidemiol. 1999;150:1094–1098. doi: 10.1093/oxfordjournals.aje.a009934. [DOI] [PubMed] [Google Scholar]
  105. Peters A, von Klot S, Heier M, Trentinaglia I, Hormann A, Wichmann HE, et al. Exposure to traffic and the onset of myocardial infarction. N Engl J Med. 2004;351:1721–1730. doi: 10.1056/NEJMoa040203. [DOI] [PubMed] [Google Scholar]
  106. Poloniecki JD, Atkinson RW, de Leon AP, Anderson HR. Daily time series for cardiovascular hospital admissions and previous day’s air pollution in London, UK. Occup Environ Med. 1997;54:535–540. doi: 10.1136/oem.54.8.535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Pope CA, III, Burnett RT, Thurston GD, Thun MJ, Calle EE, Krewski D, et al. Cardiovascular mortality and long-term exposure to particulate air pollution: epidemiological evidence of general pathophysiological pathways of disease. Circulation. 2004a;109:71–77. doi: 10.1161/01.CIR.0000108927.80044.7F. [DOI] [PubMed] [Google Scholar]
  108. Pope CA, III, Hansen ML, Long RW, Nielsen KR, Eatough NL, Wilson WE, et al. Ambient particulate air pollution, heart rate variability, and blood markers of inflammation in a panel of elderly subjects. Environ Health Perspect. 2004b;112:339–345. doi: 10.1289/ehp.6588. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Pope CA, III, Verrier RL, Lovett EG, Larson AC, Raizenne ME, Kanner RE, et al. Heart rate variability associated with particulate air pollution. Am Heart J. 1999;138:890–899. doi: 10.1016/s0002-8703(99)70014-1. [DOI] [PubMed] [Google Scholar]
  110. Pritchard RJ, Ghio AJ, Lehmann JR, Winsett DW, Tepper JS, Park P, et al. Oxidant generation and lung injury after particulate air pollutant exposure increase with the concentrations of associated metals. Inhal Toxicol. 1996;8:457–477. [Google Scholar]
  111. Pye M, Rae AP, Cobbe SM. Study of serum C-reactive protein concentration in cardiac failure. Br Heart J. 1990;63:228–230. doi: 10.1136/hrt.63.4.228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Quay JL, Reed W, Samet J, Devlin RB. Air pollution particles induce IL-6 gene expression in human airway epithelial cells via NF-kappaB activation. Am J Respir Cell Mol Biol. 1998;19:98–106. doi: 10.1165/ajrcmb.19.1.3132. [DOI] [PubMed] [Google Scholar]
  113. Rahman I, Morrison D, Donaldson K, MacNee W. Systemic oxidative stress in asthma, COPD, and smokers. Am J Respir Crit Care Med. 1996;154:1055–1060. doi: 10.1164/ajrccm.154.4.8887607. [DOI] [PubMed] [Google Scholar]
  114. Ridker PM. Role of inflammatory biomarkers in prediction of coronary heart disease. Lancet. 2001;358:946–948. doi: 10.1016/S0140-6736(01)06112-8. [DOI] [PubMed] [Google Scholar]
  115. Ridker PM, Hennekens CH, Buring JE, Rifai N. C-reactive protein and other markers of inflammation in the prediction of cardiovascular disease in women. N Engl J Med. 2000;342:836–843. doi: 10.1056/NEJM200003233421202. [DOI] [PubMed] [Google Scholar]
  116. Ridker PM, Rifai N, Lowenthal SP. Rapid reduction in C-reactive protein with cerivastatin among 785 patients with hypercholestrolemia. Circulation. 2001;103:1191–1193. doi: 10.1161/01.cir.103.9.1191. [DOI] [PubMed] [Google Scholar]
  117. Riediker M, Cascio WE, Griggs TR, Herbst MC, Bromberg PA, Neas L, et al. Particulate matter exposure in cars is associated with cardiovascular effects in healthy young men. Am J Respir Crit Care Med. 2004;169:934–940. doi: 10.1164/rccm.200310-1463OC. [DOI] [PubMed] [Google Scholar]
  118. Rifai N, Ridker PM. High-sensitivity C-reactive protein: a novel and promising marker of coronary heart disease. Clin Chem. 2001;47:403–411. [PubMed] [Google Scholar]
  119. Rossi G, Vigotti MA, Zanobetti A, Repetto F, Gianelle V, Schwartz J. Air pollution and cause-specific mortality in Milan, Italy, 1980–1989. Arch Environ Health. 1999;54:158–164. doi: 10.1080/00039899909602254. [DOI] [PubMed] [Google Scholar]
  120. Salvi S, Blomberg A, Rudell B, Kelly F, Sandströöm T, Holgate ST, et al. Acute inflammatory responses in the airways and peripheral blood after short-term exposure to diesel exhaust in healthy human volunteers. Am J Respir Crit Care Med. 1999;159:702–709. doi: 10.1164/ajrccm.159.3.9709083. [DOI] [PubMed] [Google Scholar]
  121. Samet JM, Dominici F, Curriero FC, Coursac I, Zeger SL. Fine particulate air pollution and mortality in 20 U.S. cities, 1987–1994. N Engl J Med. 2000a;343:1742–1749. doi: 10.1056/NEJM200012143432401. [DOI] [PubMed] [Google Scholar]
  122. Samet JM, Zeger SL, Dominici F, Curriero F, Cursac I, Dockery DW, et al. The National Morbidity, Mortality, and Air Pollution Study. Part II: Morbidity and mortality from air pollution in the United States. Res Rep Health Eff Inst. 2000b;94(pt 2):5–79. [PubMed] [Google Scholar]
  123. Sarnat JA, Koutrakis P, Suh HH. Assessing the relationship between personal particulate and gaseous exposures of senior citizens living in Baltimore, MD. J Air Waste Manage Assoc. 2000;50:1184–1198. doi: 10.1080/10473289.2000.10464165. [DOI] [PubMed] [Google Scholar]
  124. Sarnat JA, Schwartz J, Catalano PJ, Suh HH. Gaseous pollutants in particulate matter epidemiology: confounders or surrogates? Environ Health Perspect. 2001;109:1053–1061. doi: 10.1289/ehp.011091053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Schiffrin EL. A critical review of the role of endothelial factors in the pathogenesis of hypertension. J Cardiovasc Pharmacol. 2001;38(suppl 2):S3–S6. doi: 10.1097/00005344-200111002-00002. [DOI] [PubMed] [Google Scholar]
  126. Schreck R, Rieber P, Baeuerle PA. Reactive oxygen intermediates as apparently widely used messengers in the activation of the NF-kappa B transcription factor and HIV-1. EMBO J. 1991;10:2247–2258. doi: 10.1002/j.1460-2075.1991.tb07761.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Schwartz J. Air pollution and hospital admissions for heart disease in eight U.S. counties. Epidemiology. 1999;10:17–22. [PubMed] [Google Scholar]
  128. Schwartz J. Air pollution and blood markers of cardiovascular risk. Environ Health Perspect. 2001;109(suppl 3):405–409. doi: 10.1289/ehp.01109s3405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Schwartz J, Dockery DW, Neas LM. Is daily mortality associated specifically with fine particles? J Air Waste Manag Assoc. 1996;46:927–939. [PubMed] [Google Scholar]
  130. Schwartz J, Morris R. Air pollution and hospital admissions for cardiovascular disease in Detroit, Michigan. Am J Epidemiol. 1995;142:23–35. doi: 10.1093/oxfordjournals.aje.a117541. [DOI] [PubMed] [Google Scholar]
  131. Seaton A, MacNee W, Donaldson K, Godden D. Particulate air pollution and acute health effects. Lancet. 1995;345:176–178. doi: 10.1016/s0140-6736(95)90173-6. [DOI] [PubMed] [Google Scholar]
  132. Seaton A, Soutar A, Crawford V, Elton R, McNerlan S, Cherrie J, et al. Particulate air pollution and the blood. Thorax. 1999;54:1027–1032. doi: 10.1136/thx.54.11.1027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Shi JP, Evans DE, Khan AA, Harrison RM. Source and concentration of nanoparticles (< 10 nm diameter) in the urban atmosphere. Atmos Environ. 2001;35:1193–1202. [Google Scholar]
  134. Sioutas C, Delfino RJ, Singh M. Exposure assessment for atmospheric ultrafine particles (UFP) and implications in epidemiological research. Environ Health Perspect. 2005;113(8):947–955. doi: 10.1289/ehp.7939. [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Sloan RP, Shapiro PA, Bagiella E, Boni SM, Paik M, Bigger JT, Jr, et al. Effect of mental stress throughout the day on cardiac autonomic control. Biol Psychol. 1994;37:89–99. doi: 10.1016/0301-0511(94)90024-8. [DOI] [PubMed] [Google Scholar]
  136. Sorensen M, Daneshvar B, Hansen M, Dragsted LO, Hertel O, Knudsen L, et al. Personal PM2.5 exposure and markers of oxidative stress in blood. Environ Health Perspect. 2003;111:161–165. doi: 10.1289/ehp.111-1241344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Stec JJ, Silbershatz H, Tofler GH, Matheney TH, Sutherland P, Lipinska I, et al. Association of fibrinogen with cardiovascular risk factors and cardiovascular disease in the Framingham Offspring Population. Circulation. 2000;102:1634–1638. doi: 10.1161/01.cir.102.14.1634. [DOI] [PubMed] [Google Scholar]
  138. Sunyer J, Ballester F, Tertre AL, Atkinson R, Ayres JG, Forastiere F, et al. The association of daily sulfur dioxide air pollution levels with hospital admissions for cardiovascular diseases in Europe (the APHEA-II study) Eur Heart J. 2003;24:752–760. doi: 10.1016/s0195-668x(02)00808-4. [DOI] [PubMed] [Google Scholar]
  139. Suwa T, Hogg JC, Quinlan KB, Ohgami A, Vincent R, van Eeden SF. Particulate air pollution induces progression of atherosclerosis. J Am Coll Cardiol. 2002;39:935–942. doi: 10.1016/s0735-1097(02)01715-1. [DOI] [PubMed] [Google Scholar]
  140. Tarkiainen TH, Timonen KL, Vanninen EJ, Alm S, Hartikainen JEK, Pekkanen J. Effect of acute carbon monoxide exposure on heart rate variability in patients with coronary artery disease. Clin Physiol Funct Imaging. 2003;23:98–102. doi: 10.1046/j.1475-097x.2003.00478.x. [DOI] [PubMed] [Google Scholar]
  141. Task Force of the European Society of Cardiology and the North American Society of Pacing and Electrophysiology. Heart rate variability: standards of measurement, physiological interpretation, and clinical use. Circulation. 1996;93:1043–1065. [PubMed] [Google Scholar]
  142. Thompson SG, Kienast J, Pyke SD, Haverkate F, van de Loo JC. Hemostatic factors and the risk of myocardial infarction or sudden death in patients with angina pectoris. European Concerted Action on Thrombosis and Disabilities Angina Pectoris Study Group. N Engl J Med. 1995;332:635–641. doi: 10.1056/NEJM199503093321003. [DOI] [PubMed] [Google Scholar]
  143. Tobias HJ, Beving DE, Ziemann PJ, Sakurai H, Zuk M, McMurry PH, et al. Chemical analysis of diesel engine nanoparticles using a nano-DMA/thermal desorption particle beam mass spectrometer. Environ Science Technol. 2001;35:2233–2243. doi: 10.1021/es0016654. [DOI] [PubMed] [Google Scholar]
  144. Torre-Amione G, Kapadia S, Benedict C, Oral H, Young JB, Mann DL. Proinflammatory cytokine levels in patients with depressed left ventricular ejection fraction: a report from the Studies of Left Ventricular Dysfunction (SOLVD) J Am Coll Cardiol. 1996;27:1201–1206. doi: 10.1016/0735-1097(95)00589-7. [DOI] [PubMed] [Google Scholar]
  145. Tsutamoto T, Hisanaga T, Fukai D, Wada A, Maeda Y, Maeda K, et al. Prognostic value of plasma soluble intercellular adhesion molecule-1 and endothelin-1 concentration in patients with chronic congestive heart failure. Am J Cardiol. 1995;76:803–808. doi: 10.1016/s0002-9149(99)80231-8. [DOI] [PubMed] [Google Scholar]
  146. Urch B, Brook JR, Wasserstein D, Brook RD, Rajagopalan S, Corey P, et al. Relative contributions of PM2.5 chemical constituents to acute arterial vasoconstriction in humans. Inhal Toxicol. 2004;16:345–352. doi: 10.1080/08958370490439489. [DOI] [PubMed] [Google Scholar]
  147. U.S. EPA 2003. Air Quality Criteria for Particulate Matter, Fourth External Review Draft. EPA/600/P-99/002aD and bD. Research Triangle Park, NC:Office of Research and Development, National Center for Environmental Assessment, U.S. Environmental Protection Agency.
  148. Utell MJ, Frampton MW. Acute health effects of ambient air pollution: the ultrafine particle hypothesis. J Aerosol Med. 2000;13:355–359. doi: 10.1089/jam.2000.13.355. [DOI] [PubMed] [Google Scholar]
  149. Utell MJ, Frampton MW, Zareba W, Devlin RB, Cascio WE. Cardiovascular effects associated with air pollution: potential mechanisms and methods of testing. Inhal Toxicol. 2002;14:1231–1247. doi: 10.1080/08958370290084881. [DOI] [PubMed] [Google Scholar]
  150. van Eeden SF, Hogg JC. Systemic inflammatory response induced by particulate matter air pollution: the importance of bone-marrow stimulation. J Toxicol Environ Health A. 2002;65:1597–1613. doi: 10.1080/00984100290071685. [DOI] [PubMed] [Google Scholar]
  151. van Eeden SF, Tan WC, Suwa T, Mukae H, Terashima T, Fujii T, et al. Cytokines involved in the systemic inflammatory response induced by exposure to particulate matter air pollutants PM10. Am J Respir Crit Care Med. 2001;164:826–830. doi: 10.1164/ajrccm.164.5.2010160. [DOI] [PubMed] [Google Scholar]
  152. Vanhoutte PM. Say NO to ET. J Auton Nerv Syst. 2000;81:271–277. doi: 10.1016/s0165-1838(00)00126-0. [DOI] [PubMed] [Google Scholar]
  153. Venners SA, Wang B, Xu Z, Schlatter Y, Wang L, Xu X. Particulate matter, sulfur dioxide, and daily mortality in Chongqing, China. Environ Health Perspect. 2003;111:562–567. doi: 10.1289/ehp.5664. [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Veronesi B, Oortgiesen M. Neurogenic inflammation and particulate matter (PM) air pollutants. Neurotoxicology. 2001;22:795–810. doi: 10.1016/s0161-813x(01)00062-6. [DOI] [PubMed] [Google Scholar]
  155. Wedzicha JA, Seemungal TA, MacCallum PK, Paul EA, Donaldson GC, Bhowmik A, et al. Acute exacerbations of chronic obstructive pulmonary disease are accompanied by elevations of plasma fibrinogen and serum IL-6 levels. Thromb Haemost. 2000;84:210–215. [PubMed] [Google Scholar]
  156. Weisensee D, Bereiter-Hahn J, Schoeppe W, Low-Friedrich I. Effects of cytokines on the contractility of cultured cardiac myocytes. Int J Immunopharmacol. 1993;15:581–587. doi: 10.1016/0192-0561(93)90075-a. [DOI] [PubMed] [Google Scholar]
  157. Wichmann HE, Spix C, Tuch T, Wolke G, Peters A, Heinrich J, et al. Daily mortality and fine and ultrafine particles in Erfurt, Germany. Part I: role of particle number and particle mass. Res Rep Health Eff Inst. 2000;98:5–86. [PubMed] [Google Scholar]
  158. Woods A, Brull DJ, Humphries SE, Montgomery HE. Genetics of inflammation and risk of coronary artery disease: the central role of interleukin-6. Eur Heart J. 2000;21:1574–1583. doi: 10.1053/euhj.1999.2207. [DOI] [PubMed] [Google Scholar]
  159. Yeates DB 2000. Neurally mediated cardiopulomonary and systemic responses to inhaled irritants and antigens. In: Particle-Lung Interactions (Gehr P, Heyder J, eds). New York:Marcel Dekker, 603–626.
  160. Yokoyama T, Vaca L, Rossen RD, Durante W, Hazarika P, Mann DL. Cellular basis for the negative inotropic effects of tumor necrosis factor-alpha in the adult mammalian heart. J Clin Invest. 1993;92:2303–2312. doi: 10.1172/JCI116834. [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Zanobetti A, Canner MJ, Stone PH, Schwartz J, Sher D, Eagan-Bengston E, et al. Ambient pollution and blood pressure in cardiac rehabilitation patients. Circulation. 2004;110:2184–2189. doi: 10.1161/01.CIR.0000143831.33243.D8. [DOI] [PubMed] [Google Scholar]
  162. Zanobetti A, Schwartz J. Are diabetics more susceptible to the health effects of airborne particles? Am J Respir Crit Care Med. 2001;164:831–833. doi: 10.1164/ajrccm.164.5.2012039. [DOI] [PubMed] [Google Scholar]
  163. Zanobetti A, Schwartz J, Dockery DW. Airborne particles are a risk factor for hospital admissions for heart and lung disease. Environ Health Perspect. 2000b;108:1071–1077. doi: 10.1289/ehp.001081071. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Zanobetti A, Schwartz J, Gold D. Are there sensitive subgroups for the effects of airborne particles? Environ Health Perspect. 2000a;108:841–845. doi: 10.1289/ehp.00108841. [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Zanobetti A, Schwartz J, Samoli E, Gryparis A, Touloumi G, Peacock J, et al. The temporal pattern of respiratory and heart disease mortality in response to air pollution. Environ Health Perspect. 2003;111:1188–1193. doi: 10.1289/ehp.5712. [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Zhu Y, Hinds WC, Kim S, Sioutas C. Concentration and size distribution of ultrafine particles near a major highway. J Air Waste Manage Assoc. 2002;52:1032–1042. doi: 10.1080/10473289.2002.10470842. [DOI] [PubMed] [Google Scholar]

Articles from Environmental Health Perspectives are provided here courtesy of National Institute of Environmental Health Sciences

RESOURCES